a 1,3-dipolar cycloaddition approach to the synthesis...

93
A 1,3-DIPOLAR CYCLOADDITION APPROACH TO THE SYNTHESIS OF RESINIFERATOXIN by Jennifer A. Loyer-Drew B.S., Western Washington University, 2003 Submitted to the Graduate Faculty of Arts and Sciences in partial fulfillment of the requirements for the degree of Master of Science University of Pittsburgh 2008

Upload: phungnguyet

Post on 16-Sep-2018

229 views

Category:

Documents


1 download

TRANSCRIPT

A 1,3-DIPOLAR CYCLOADDITION APPROACH TO THE SYNTHESIS OF RESINIFERATOXIN

by

Jennifer A. Loyer-Drew

B.S., Western Washington University, 2003

Submitted to the Graduate Faculty of

Arts and Sciences in partial fulfillment

of the requirements for the degree of

Master of Science

University of Pittsburgh

2008

UNIVERSITY OF PITTSBURGH

ARTS AND SCIENCES

This thesis was presented

by

Jennifer A. Loyer-Drew

It was defended on

July 31, 2008

and approved by

Professor Tara Y. Meyer

Professor Dennis P. Curran

Thesis Advisor: Professor Kay M. Brummond

ii

A 1,3-DIPOLAR CYCLOADDITION APPROACH TO THE SYNTHESIS OF

RESINIFERATOXIN

Jennifer A. Loyer-Drew, M.S.

University of Pittsburgh, 2008

The Rh(I)-catalyzed allenic cyclocarbonylation reaction is a formal [2 + 2 + 1] cycloaddition

process that has been used to gain access to 4-alkylidenecyclopentenones. Incorporation of a

six-membered ring on the tether between the allene and the alkyne components allows access to

a variety of [6-7-5] ring structures featured in the skeletons of various natural products, including

resiniferatoxin. This thesis describes the development of two systems, each with a future

synthesis of resiniferatoxin in mind. First, a model system was designed to demonstrate the

compatibility of the isoxazoline moiety with the Rh(I)-catalyzed cyclocarbonylation reaction.

The second investigation involved the synthesis of an asymmetrically functionalized 2-

cyclohexenone in order to attempt a stereoselective 1,3-dipolar cycloaddition. The first model

system successfully led to the synthesis of the unfunctionalized [6-7-5] core of resiniferatoxin

via cyclocarbonylation of an isoxazoline-containing allene-yne. Unfortunately, under numerous

conditions, the functionalized cyclohexenone synthesized for the second study failed to undergo

1,3-dipolar cycloaddition with a nitrile oxide.

iii

TABLE OF CONTENTS

ABBREVIATIONS ....................................................................................................................... x

1.0 INTRODUCTION................................................................................................................ 1

1.1 RESINIFERATOXIN ................................................................................................. 1

1.2 STRUCTURALLY-RELATED NATURAL PRODUCTS ...................................... 5

2.0 PREVIOUS APPROACHES TO THE SYNTHESIS OF RESINIFERATOXIN AND

RELATED COMPOUNDS .......................................................................................................... 7

3.0 RETROSYNTHETIC ANALYSIS OF RESINIFERATOXIN ..................................... 11

3.1 ACCESSING THE A AND B RINGS VIA A Rh(I)-CATALYZED ALLENIC

CYCLOCARBONYLATION REACTION ..................................................................... 11

3.2 THE MASKED ALDOL STRATEGY: 1,3-DIPOLAR CYCLOADDITION TO

FORM AN ISOXAZOLINE .............................................................................................. 13

4.0 RESULTS AND DISCUSSION ........................................................................................ 18

4.1 INTRODUCTION ..................................................................................................... 18

4.2 MODEL SYSTEM CONTAINING AN UNFUNCTIONALIZED C RING ........ 19

4.3 SYNTHESIS OF AN ASYMMETRICALLY-FUNCTIONALIZED C RING .... 26

4.3.1 Conjugate addition to cyclohexadienone 62 to form enone 60 ................... 27

4.3.2 Elaboration of enone 60 ................................................................................. 32

5.0 CONCLUSION .................................................................................................................. 34

iv

6.0 EXPERIMENTAL ............................................................................................................. 35

6.1 GENERAL ................................................................................................................. 35

6.2 SYNTHESIS AND CHARACTERIZATION ......................................................... 36

APPENDIX A .............................................................................................................................. 49

BIBLIOGRAPHY ....................................................................................................................... 74

v

LIST OF TABLES

Table 1. Magnesium alkoxide direction in 1,3-dipolar cycloaddition (Kanemasa) ..................... 15

Table 2. Hydrogen bond direction in 1,3-dipolar cycloaddition (Choi) ....................................... 15

Table 3. Propargylation of aldehyde 50 ........................................................................................ 24

Table 4. Literature examples of conjugate addition to quinone monoketal 62 ............................. 28

Table 5. Product distribution in copper-catalyzed conjugate addition to cyclohexadienone 62 .. 30

Table 6. Attempted stereoselective cycloaddition of enone 26 .................................................... 33

vi

LIST OF FIGURES

Figure 1. Structures of resiniferatoxin and capsaicin ...................................................................... 1

Figure 2. Phorbol and prostratin .................................................................................................... 4

Figure 3. Daphnane natural products .............................................................................................. 6

Figure 4. Molecular modeling of cyclocarbonylation products .................................................... 17

Figure 5. 1H NMR verification of regioselectivity ...................................................................... 21

Figure 6. Minimum energy conformations and dihedral angles for 59α (left) and 59β (right) ... 26

Figure 7. Chiral phosphoramidite ligands ..................................................................................... 28

vii

LIST OF SCHEMES

Scheme 1. Wender's synthesis of resiniferatoxin ............................................................................ 8

Scheme 2. Phenol p-alkylation to yield spirocycle 11 .................................................................... 9

Scheme 3. Carreira's approach to resiniferatoxin ........................................................................... 9

Scheme 4. Cha's synthesis of phorbol ........................................................................................... 10

Scheme 5. Cyclocarbonylation of a tethered allene-yne ............................................................... 11

Scheme 6. Access of [6-7-5] skeletons via allenic cyclocarbonylation reaction ......................... 12

Scheme 7. Retrosynthetic analysis of resiniferatoxin ................................................................... 13

Scheme 8. Steric direction of nitrile oxide cycloaddition (Martin) .............................................. 16

Scheme 9. Synthesis of model system ......................................................................................... 19

Scheme 10. Dependence of chemical shift on regioselectivity in nitrile oxide cycloaddition

(Grünanger) ................................................................................................................................... 20

Scheme 11. Propargylation of aldehyde 50 .................................................................................. 22

Scheme 12. Regioselectivity in propargylation reactions ............................................................. 23

Scheme 13. Cyclocarbonylation of the model system .................................................................. 25

Scheme 14. Proposed synthesis of functionalized C ring ............................................................. 27

Scheme 15. Synthesis of racemic phosphoramidite ligand 63 ...................................................... 29

Scheme 16. Improved synthesis of phosphoramidite (S,R,R)-64 ................................................. 31

Scheme 17. Synthesis of enone 28 ............................................................................................... 32

viii

Thank you, Erech. Go Team Venture!

ix

x

ABBREVIATIONS

Ac2O Acetic anhydride

9-BBN 9-Borabicyclo[3.3.1]nonane

BINOL 1,1’-Bi(2-naphthol)

DCM Dichloromethane

DIPA Diisopropylamine

DMAP 4-N,N-Dimethylaminopyridine

DMF N,N-Dimethylformamide

DMSO Dimethylsulfoxide

EtOAc Ethyl acetate

NMR Nuclear magnetic resonance

PAA para-Anisaldehyde

PhH Benzene

TBS tert-Butyldimethylsilyl

TEA Triethylamine

THF Tetrahydrofuran

TLC Thin layer chromatography

TMEDA N,N,N’,N’-Tetramethylethylenediamine

TMS Trimethylsilyl

1.0 INTRODUCTION

1.1 RESINIFERATOXIN

The therapeutic benefits of euphorbium—the dried latex of plants of the genus Euphorbia—have

been appreciated for several centuries; in his posthumously published Tractatus de Materia

Medica, 18th century French chemist and physician Etiénne-François Geoffroy cites euphorbium

as an effective remedy for bone cavities and nerve pains.1 Although euphorbium disappeared

from the documented pharmacopoeia in the 1800s, interest in its active constituent,

resiniferatoxin (1, Figure 1), has been renewed, as resiniferatoxin shows potential in treating,

among other ailments, chronic pain and bladder incontinence.

OO

O9

13

41

OOH

H

Ph

20

OO

OH

OMe

1

AB

COMe8

NHOH

O

( )4

2

Figure 1. Structures of resiniferatoxin and capsaicin

Resiniferatoxin was isolated in 1975 from the latex of Euphorbia resinifera and related

plants by Hecker and colleagues.2 The high irritant activity of the compound was immediately

recognized as novel and, seven years later, Hecker published a revised structure of the natural

product along with the results of preliminary structure-activity studies.3 Although the

1

pharmacophore has not been determined, the phenol and orthoester moieties have been deemed

essential for potency.

Resiniferatoxin is a known agonist of the transient receptor potential vanilloid (TRPV1).

Analogous physiological effects and structural similarities to capsaicin (2)—also an irritant that

binds to a TRPV1—suggest a common mode of action for the two compounds. Both natural

products induce pain, neurogenic edema and hypothermia in rats.4

Vanilloid receptors are ligand-gated cation channels located in the membranes of

nociceptive sensory nerves. Activation of the proteins by chemical agonists or heat causes the

channel to open, leading to an influx of intracellular calcium cations; this depolarization

generates an action potential that is perceived by the central nervous system and leads to a

burning sensation. In 1997, Julius and coworkers5 cloned TRPV1 and characterized the protein

as containing six transmembrane domains. The vanilloid binding pocket was later located

through the use of radiolabeled resiniferatoxin, and was shown to exist between domains three

and four.6

Although capsaicin and resiniferatoxin cause initial irritation at the site of application,

tachyphylaxis is demonstrated upon subsequent exposure. This phenomenon is much more

prominant with the use of resiniferatoxin than with capsaicin, although the initial irritation

caused by resiniferatoxin is only marginally more severe. The exact mechanism of

desensitization remains elusive, but summaries of thoughtful speculation can be found in several

reviews.7-9

Not only are nociceptors exposed to resiniferatoxin densensitized to further treatment with

resiniferatoxin, but they often show diminished responses to other stimuli as well, including

capsaicin, heat and other exogenous inflammatory agents. Thus, resiniferatoxin is viewed as a

2

potentially valuable analgesic for those suffering from chronic neurogenic pain and other

phenomena that result from hyperactive nociceptors.

In particular, some subjects exhibiting symptoms of overactive bladder and incontinence

have been reported to have an abnormally high density of nociceptive neurons present in their

bladder tissues. Introduced intravesically, low concentration solutions of resiniferatoxin have

been shown to lead to a decreased frequency in incontinent episodes in some of these patients;

the effects of a single treatment can last up to three months and the initial discomfort is

minimal.10, 11

Current investigation of clinical applications of resiniferatoxin seems to be taking place

predominantly in the academic arena. This may be due, in part, to a 2004 press release issued by

ICOS stating their findings that resiniferatoxin did not pass Phase II trials for treatment of

interstitial cystitis12 and that ICOS’s interest in resiniferatoxin had ceased.13 Since this statement

by ICOS, however, additional contradictory reports14 demonstrating the efficacy of

resiniferatoxin in treating bladder pain syndrome and interstitial cystitis have appeared in the

literature.

Some scientists have expressed reservations with respect to treating humans with

resiniferatoxin based on the similarities between its structure and those of the tumor-promoting

phorbol esters (12,13-diesters of 3, R1 = OH, R2 = H, Figure 2);15 however, these concerns are

not corroborated by experimental evidence. Notably, prostratin (3, R1 = H, R2 = Ac, Figure 2), a

12-deoxytigliane recently synthesized by Wender and coworkers16 from phorbol, was shown to

be a non-tumor-promoting potential anti-HIV therapeutic. Prostratin is currently in preclinical

development.

3

12 13H

R2OR1

OHH

OHO

OH

3phorbol R1 = OH, R2 = H

prostratin R1 = H, R2 = Ac

Figure 2. Phorbol and prostratin

The FDA designated resiniferatoxin as an orphan drug in 2003 for the treatment of

“intractable pain at end-stage disease.”17 This classification seems especially befitting in light of

a 2005 publication from the School of Veterinary Medicine at the University of Pennsylvania.

Administered intrathecally, resiniferatoxin appeared to largely diminish the pain associated with

bone cancer as experienced by a group of canine companion animals.18 When the dogs entered

the study, many were not deemed to be achieving adequate pain relief through the use of

conventional analgesics. After treatment with resiniferatoxin, the comfort level of most of the

dogs seemed to improve so drastically that analgesic use was tapered or in some cases

discontinued completely. No lasting ill effects of resiniferatoxin were observed, pre- or

postmortem.

TRPV1 has become a popular target for development of analgesics; a number of TRPV1

antagonists are in clinical development to address painful phenomena ranging from migraine

headaches to HIV neuropathy-associated pain.19, 20 Capsaicin is available over-the-counter as a

topical ointment for the treatment of arthritis pain. Recent publications regarding the therapeutic

possibilities of resiniferatoxin include a study of its use as a long-lasting local anasthetic21 and a

patent application alleging the efficacy of injections of resiniferatoxin in treating joint pain.22

4

The isolation of resiniferatoxin from euphorbium continues to be a materials-intensive

and fairly noxious process. Fattorusso23 has published an improved procedure for the isolation

of resiniferatoxin, but only managed a 0.0020% yield from the fresh latex of the E. resinifera

plant. One of the altruistic goals of the synthesis of resiniferatoxin is to provide a vehicle for

synthesis and testing of analogs that are more accessible or that perhaps possess more optimal

biological activity than the natural product itself.

1.2 STRUCTURALLY-RELATED NATURAL PRODUCTS

While it is our objective to carry out a synthesis of resiniferatoxin, our expectation is that the

methods that we utilize will find application in the synthesis of structurally-similar molecules.

Many daphnane diterpenoids have been isolated from natural sources; they collectively exhibit a

wide variety of fascinating and potentially useful biological activities.24 Several of these natural

products are depicted in Figure 3 with their corresponding biological activities.25-31 Phorbol (3,

R1 = OH, R2 = H, Figure 2), a tigliane diterpene, is also notable not only for the biological

activity of its 12,13-diesters, but also for the inspirational amount of creative chemistry that has

been developed in pursuit of its synthesis.

5

OO

OH

OH

Gnidimacrin26, 27

anti-tumor agent, PKC activator

O

OHOHOBz

HO

( )4

R

OO

O

OOH

H

OH

Maprouneacin28

exhibits antihyperglycemic activity

O

O

O

O

O

O

( )3R =

Ph

OO

OH

OH

Rediocide A30

potent anti-flea compound

OH

O

OHOH

H

O

O

O

O

Ph

OO

OH

Genkwanine D31

inhibitor of endothelium cell proliferationand cytotoxic towards tumor cells

OHOHOBz

OHOH

OH

OO

OH

OH

Kirkinine B29

possesses neutrophic and antitumor activity

O

OHOH

( )8

O

OO

OH

OH

Huratoxin25

potent piscicidal activity

O

OHOHO

( )7

C

AB

Daphnane skeleton

Figure 3. Daphnane natural products

6

2.0 PREVIOUS APPROACHES TO THE SYNTHESIS OF RESINIFERATOXIN AND

RELATED COMPOUNDS

To date, there has been only one total synthesis of resiniferatoxin, published by Wender in

1997.32 The absolute stereochemistry was set in the first step: known epoxide 4 (Scheme 1) was

synthesized from 1,4-pentadien-3-ol using Sharpless’ asymmetric epoxidation conditions.

Subsequent steps proceeded stereoselectively, affording intermediate 5. The key step in

Wender’s synthesis was a [5 + 2] intramolecular cycloaddition of oxidopyrylium 6, forming the

B and C rings simultaneously. An advantage of this tactic was the resulting rigidity of the

tricyclic ring system. The presence of the bridging ether in 7 lent bias to the formation of

subsequent stereocenters. A zirconium-mediated cyclization of enyne 8 closed the A ring of

resiniferatoxin and further elaboration of 9 led to completion of the natural product.

Esterification at C20 and construction of the orthoester were performed towards the end of the

synthesis, since these moieties have been shown to be essential for potent irritant activity.

7

OH

O

OBn

OAcO

O

OAcOTBS

DBU

MeCN, 80 °C

OAc

OBnO

O

OTBS

OAc

OBn

OO

OTBS

84%

6

4 5

steps

7

OAc

OBn

OOTBS

8

OTMSPh

Cp2ZrBu2, THF, -78 °C

then HOAc, 90%

OH

OBn

OOTBS

9

OTMSPh

steps

steps1

H

Scheme 1. Wender's synthesis of resiniferatoxin

Carreira’s group33, 34 has published two papers disseminating complementary approaches

to the construction of angular [6-7-5] ring structures. Substrates and conditions for

diastereoselective phenol para-alkylation were developed in order to gain access to spiro-fused

cyclohexadienones (Scheme 2).34 Conversion of optically pure alcohol 10 to the 3,5-

bistrifluoromethylbenzenesulfonate ester and subsequent Winstein alkylation provided

cyclohexadienone 11 as a single stereoisomer in 95% yield. Previously, Carreira and

colleagues33 had shown that tricyclic compound 12 could undergo photorearrangement via

intermediate 13 to directly provide the [6-7-5] core 14 of resiniferatoxin and similar natural

products (Scheme 3). Enone 14 was elaborated further to give the highly oxygenated ketal 14 in

order to demonstrate the synthetic potential of the photorearrangement approach.

8

OEtBzO

OO

H

PhOH

OBz F3C

CF3

SO2Cl

DMAP, DCM

i,

ii, K2CO3, EtOH/i-PrOH95% OEt

O

O

O PhHO

H

10 11

Scheme 2. Phenol p-alkylation to yield spirocycle 11

OEtO

OMe

HOH

OH

OH

OHH

OEtO

OMe

TFA/pentane

O

EtO

O OHH

HOMe

82%

HH

OMe

OHO

HO

OHMeO OMe

12

14 15

13

Scheme 3. Carreira's approach to resiniferatoxin

The Wender group has applied their [5 + 2] cycloaddition chemistry several times to the

synthesis of phorbol; their most recent synthesis was asymmetric.35-38 Wender39 recently

published preliminary results on the application of this methodology towards the synthesis of

gnidimacrin.

The only other complete synthesis of phorbol to date was published by Lee and Cha40 in

2001. Like Wender, Lee and Cha employed a cycloaddition as a key step. [4 + 3] Cycloaddition

of furan 16 (Scheme 4) and the oxyallyl generated from 1,1,3-trichloroacetone (17), followed by

dechlorination with zinc, afforded meso compound 18 (R = TBS). Desymmeterization of the

corresponding acetate (18, R = Ac) was performed enzymatically to give alcohol 19 in 90% yield

9

and 80% ee. Subsequent steps proceeded diastereoselectively to give substrate 20 which

underwent an intramolecular Heck reaction to close the C ring. Lee and Cha later intersected

with Wender’s synthesis of phorbol.

There have been numerous innovative approaches to the construction of a suitably

functionalized angular [6-7-5] tricyclic core of phorbol; these strategies include: Diels–Alder

cyclization,41-43 an intramolecular 1,3-dipolar cycloaddition,44, 45 a diyl-trapping cycloaddition,46

and an anionic oxy-Cope rearrangement.47 The obvious challenges of constructing the polycyclic

core of resiniferatoxin and related natural products that possess opportunely located functionality

have inspired many chemists in the synthetic community, including the Brummond group.

O

OTBS

OTBS

Cl

Cl

OCl

OR

OR

O O

OAc

OH

O Oi, TEA, TFE

ii, Zn, MeOH+

Candida rugosa

OTBS

O

I

3

16 17 18 19

20

TMSO

Ph

steps steps

Scheme 4. Cha's synthesis of phorbol

10

3.0 RETROSYNTHETIC ANALYSIS OF RESINIFERATOXIN

3.1 ACCESSING THE A AND B RINGS VIA A Rh(I)-CATALYZED ALLENIC

CYCLOCARBONYLATION REACTION

The Brummond group has been preeminent in applying transition metal catalysis to the creation

of a diverse array of polycyclic scaffolds via the Pauson–Khand-type cyclocarbonylation of

allenes.48-55 Former group members have demonstrated the utility of the molybdenum-mediated

and rhodium-catalyzed methodologies in their application to natural product synthesis.48, 50-52 A

unique feature of the rhodium-catalyzed system is its selectivity for reaction with the distal

double bond of the allene;49 in this way, our group has been able to selectively access a variety of

4-alkylidene cyclopentenones by varying the length of the tether between the allene and alkyne

moieties (Scheme 5).

•O

[Rh(CO)2Cl]2, CO

Scheme 5. Cyclocarbonylation of a tethered allene-yne

Incorporation of a six-membered ring on the tether between the allene and alkyne has

allowed access to a variety of angular and linear [6-7-5] tricyclic skeletons present in several

natural products (Scheme 6).48, 50 The angular [6-7-5] substructure present in resiniferatoxin

seems specially suited to showcase our group’s cyclocarbonylation methodology.48 The

11

unsaturation present in the cyclocarbonylation product is conveniently situated at sites requiring

oxidation. Further, the rhodium-catalyzed reaction has demonstrated excellent functional group

compatibility. In the case of resiniferatoxin, this tolerance would permit us to submit an allene-

yne tethered by a functionalized C ring to the cyclocarbonylation reaction, yielding an advanced

synthetic intermediate.

HO

R

( )n

( )m

R

O

R

O

R

O

RO

OO

O

OOH

H

Ph

OR

1

OHC

OAc

O

HO HOOH

HOHO OH

H H

H

OH

Guanacastepene A Grayanotoxin III Rippertene

Scheme 6. Access of [6-7-5] skeletons via allenic cyclocarbonylation reaction

We envision a formal synthesis of resiniferatoxin, intersecting Wender’s total synthesis at

intermediate alcohol 21 (Scheme 7). We intend to gain access to this intermediate via functional

group manipulation of 22. A key premise of our proposal is the use of the isoxazoline moiety as

a masked aldol equivalent (see section 3.2); thus, opening of the isoxazoline ring in 23 will yield

β-hydroxy ketone 22. Isoxazoline 23 may be obtained by performing selective oxidations and

12

olefination of 24, the cyclocarbonylation product of allenyne 25. Conversion of the ketone

present in 26 to the allene, followed by deprotection of the TBS ether, oxidation to the aldehyde,

and propargylation will yield allene-yne 25. We propose that a regio- and stereospecific 1,3-

dipolar cycloaddition of enone 28 and nitrile oxide 27 will provide access to isoxazoline 26.

OO

O OH

OOH

H

Ph

21

O

OPOP

OO

OPH

OP

OO

OPHN

O

22 23

OP

OP

O

NO

NO

OP

R

OP

O

OP

NO

OTBSO

OHON

OTBS

24

+

25262728

PO PO

PO PO PO

PO

Scheme 7. Retrosynthetic analysis of resiniferatoxin

3.2 THE MASKED ALDOL STRATEGY: 1,3-DIPOLAR CYCLOADDITION TO

FORM AN ISOXAZOLINE

1,3-Dipolar cycloaddition of nitrile oxides to provide isoxazolines followed by reduction of the

N-O bond and hydrolysis of the resulting imine is a strategy for accessing β-hydroxy ketones.56

While numerous synthetic innovations allow for control of aldol product stereochemistry and

preferential formation of a singular cross-aldol product, the cycloaddition of nitrile oxides and

13

olefins continues to offer an elegant alternative to the aldol reaction.57-61 The characteristics that

attracted us to the nitrile oxide cycloaddition approach were:

(1) The potential for selectively establishing two stereocenters and latent

functionality in a single step;

(2) The ability of the isoxazoline to act as a protecting group for functionality to be

unmasked late in the synthesis; and

(3) Rigidification of the cyclocarbonylation product to provide a clear facial

preference for subsequent oxidations.

We anticipate that the regioselectivity of the cycloaddition will be influenced by the

presence of the enone. High regioselectivity has been reported for the reaction of nitrile oxides

with cycloalkenones, largely due to steric effects.62 We anticipate that the hydroxyl group may

aid in guiding the nitrile oxide to a syn approach, as well as in reinforcing the regioselectivity;

although, this type of assistance is not well supported in the literature. While hydrogen-bonding

has not been observed to be a strong enough force to consistently direct the 1,3-dipolar

cycloaddition of nitrile oxides,63-65 supplanting the proton of the alcohol with a stronger Lewis

acid has been demonstrated to lead to high regio- and stereoselectivity.66-70

Kanemasa and colleagues70 found that conversion of allylic alcohols to the corresponding

magnesium alkoxide (e.g., 30, X = MgBr, Table 1) and subsequent exposure to benzonitrile

oxide (generated through dehydrohalogenation of hydroximoyl chloride 29) gave the

corresponding isoxazolines in a highly regio- and stereoselective fashion (Table 1, entry 2). It

was noted, however, that the reaction of the magnesium alkoxide generated from 2-cyclohexenol

“led to the formation of a complex mixture of many products.”

14

Table 1. Magnesium alkoxide direction in 1,3-dipolar cycloaddition (Kanemasa)

NOH

ClPh

baseOX

Ph

N O

OHPh

N O

OHPh

N O

OH

+ +

31 32 3329

30

Base X Yield (%) 31 : 32 : 33

TEA H 28 55 : 27 : 18

EtMgBr MgBr 53 94 : 6 : 0

A related result was realized in a study of the directing abilities of 2° amides conducted

by a member of the Curran group. After exposure to conditions that led to satisfactory formation

of the syn-cycloadduct 36 of the corresponding cyclopentenyl amide (35, n = 0, Table 2, entry 1),

unreacted cyclohexenyl amide (35, n = 1) was recovered, even when the reaction was heated to

80 °C for nearly 10 days (Table 2, entries 2 and 3).68 Choi suggested this result was due to the

relatively low reactivity of cyclohexene-derived substrates towards cycloaddition and to the

inability of such substrates to adopt the necessary hydrogen-bonded transition state.

Table 2. Hydrogen bond direction in 1,3-dipolar cycloaddition (Choi)

N Ph

OH

TEA N Ph

OH

ON

t-Bu

+NOH

Clt-Bu

N Ph

OH

ON

t-Bu

N Ph

OH

NO

t-Bu+ +

34 35 36 37 38

( )n ( )n ( )n ( )n

Entry n Time (days) T (°C) Yield (%) 36 : 37 : 38

1 0 4.5 25 92 85 : 1 : 14

2 1 4.5 25 NR NA

3 1 9.5 80 NR NA

15

It is possible that the facial selectivity of the 1,3-dipolar cycloaddition will be controlled

by sterics alone. In studies directed toward the synthesis of breynolide, the Martin group71, 72

observed a single cycloadduct 41 resulting from addition of the nitrile oxide generated from

hydroximoyl chloride 40 anti to a large substituent on a cyclohexenone substrate 39 (Scheme 8).

O

ON

SBnO

H

H

O

SBnO

H

H

+

NOH

ClO O

TEA, Et2O

37%

O

O

39 40 41

Scheme 8. Steric direction of nitrile oxide cycloaddition (Martin)

The presence of the isoxazoline ring will impose considerable conformational constraints

on our synthetic intermediates. Molecular modeling of the two possible diastereomers (resulting

from the propargylation step) of the cyclocarbonylation product using Spartan shows that each

molecule possesses a convex and a concave face (Figure 4). In both cases, the more sterically

accessible convex face is the face on which we hope to perform selective oxidations of the

enone.

16

O

TMSO

NO OMe

O

TMSO

NO OMe

Figure 4. Molecular modeling of cyclocarbonylation products

17

4.0 RESULTS AND DISCUSSION

4.1 INTRODUCTION

Prior to committing to the synthesis of resiniferatoxin, our retrosynthetic plan required that we

address three questions:

(1) Is the isoxazoline moiety tolerated by the reaction conditions we wish to

employ, namely the Rh(I)-catalyzed cyclocarbonylation protocol?

(2) Is selective oxidation of the resulting alkylidene cyclopentenone feasible?

(3) Are we able to carry out a stereoselective [3 + 2] cycloaddition?

In order to satisfy these queries, we embarked on two separate investigations. We utilized a

model system to demonstrate the compatibility of the isoxazoline moiety with the rhodium-

catalyzed cyclocarbonylation reaction (section 4.2). Also, this model system will allow us to

ascertain the utility of the cyclocarbonylation product in subsequent selective oxidation.

Concurrently, an asymmetrically functionalized C ring was prepared to probe the possibility of

obtaining selectivity in the 1,3-dipolar cycloaddition (section 4.3).

18

4.2 MODEL SYSTEM CONTAINING AN UNFUNCTIONALIZED C RING

OHO2N

TBSCl, imidazole

DMFOTBS

O2N

43

OCN

NCO

TEA, PhH, refluxO

NO

OTBS45

NO

OTBS

46

THF, 0 °C

MgBr

HO DCM, 0 °C

NO

OTBS

47

AcO

Ac2O, TEA, DMAP

toluene, H2O

[(Ph3P)CuH]6

NO

OTBS

48

•THF, 0 °C

TBAF NO

OH

49

•TEA, DCM, -65 °C

(COCl)2, DMSO NO

O

50

92%

2-cyclohexenone (44)

43%

84% 74% 87%

89% 75%

42

Scheme 9. Synthesis of model system

Under my direction, undergraduate researcher Darla Seifried explored the steps shown in

Scheme 9.73 I have since repeated the reactions several times to investigate the propargylation

and subsequent cyclocarbonylation steps. All yields reported are from my syntheses.

Nitroethanol (42) was protected as the TBS ether 43, which then underwent in situ dehydration

and 1,3-dipolar cycloaddition with 2-cyclohexenone (44) to give bicycle 45 as a single

regioisomer. Initially, phenylisocyanate was used as the dehydrating agent;74 however, the

diphenyl urea byproduct generated during the reaction was extremely difficult to remove from

the desired product via either multiple precipitations (using ethyl acetate, hexanes, ether, or

water) of the urea or column chromatography.73 Kurth and colleagues75 have reported a clever

solution to this frequently observed problem: the use of 1,4-phenylene diisocyanate generates a

19

polyurea that is insoluble in organic solvents (THF, methylene chloride and benzene) and can

therefore be removed by filtration. We found Kurth’s method to be satisfactory, although a

single filtration is rarely sufficient to remove the significant quantities of polymeric urea present

in the reaction mixture.

The regioselectivity of the cycloaddition was confirmed by 1H NMR. In isoxazoline 45,

H5 gives a doublet of triplets resonance at δ 4.95 ppm and H4 appears as a doublet at δ 3.86 ppm

(Figure 5). This is consistent with the findings of Grünanger and colleagues62 in their studies of

the relationship between regioisomers and chemical shift in similar systems. In the addition of

benzonitrile oxide to 2-cyclohexenone (44), the multiplet resonance for H5 appears significantly

downfield from the H4 doublet in the major product 51 (Scheme 10). The ΔδH4,H5 is much

smaller for the regioisomer 52, and the more deshielded proton—H5—is alpha to the ketone and

therefore only split by H4 (i.e., appears as a doublet). Grünanger observed only cis-addition to

cycloalkenones, the products exhibiting a characteristically large 3JH4,H5 (8.8 to 12.0 Hz). The 3J

coupling (9.7 Hz) exhibited by H4 and H5 in isoxazoline 45 falls neatly within this range.76, 77

45 N

O

Ph

H

H

O

OPh

NOH

Cl

TEA

Et2O, rt, 12 h85%

3 : 1 (51 : 52)

+ +

2944 51H4: δ 4.23 (d)H5: δ 5.12 (m)

52H4: δ 4.20 (m)H5: δ 4.68 (d)

45 N

O

Ph

H

HO

Scheme 10. Dependence of chemical shift on regioselectivity in nitrile oxide cycloaddition (Grünanger)

20

45 N

OH

HO

45OTBS

H4

H5

Figure 5. 1H NMR verification of regioselectivity

Addition of ethynylmagnesium bromide to the ketone 45, followed by acylation of the

tertiary alcohol 46, gave propargyl acetate 47 as a single diastereomer. As has been experienced

by other group members,48, 51 Stryker’s reagent ([(Ph3P)CuH]6) was suitable for the reduction of

the propargyl acetate 47 to the corresponding 1,1-disubstituted allene 48. Initial yields obtained

for the reduction of the propargyl acetate 47 using commercial samples of Stryker’s reagent were

around 45%. Others78-80 have remarked on the variable (and generally inferior) quality of

commercial [(Ph3P)CuH]681 as compared to Stryker’s reagent that has been prepared in the

laboratory. Consequently, the copper hydride species was prepared according to the literature82

and, using the fresh reagent, the yield of allene 48 nearly doubled to 87%. Deprotection of the

TBS ether 48 followed by Swern oxidation83 of the primary alcohol 49 gave aldehyde 50.

21

The addition of nucleophiles to 3-formyl-Δ2-isoxazolines is precedented,84 although

propargylation has not been reported. A zinc Barbier reaction of the aldehyde 50 and propargyl

bromide in aqueous ammonium chloride produced the desired homopropargyl alcohol 53 in

variable yield as a 3:1 mixture of diastereomers (Scheme 11). However, subjecting this substrate

to the [Rh(CO)2Cl]2 cyclocarbonylation conditions led mostly to decomposition. Other group

members have observed similar behavior in the case of terminal alkynes.48 Attempts at protecting

the terminal alkyne 53 with a trimethylsilyl group proved unsuccessful and efforts to force the

formation of 54 by exposure of the protected alcohol to excess base led primarily to

decomposition.

NO

O•

Br

Zn, sat NH4Cl, THF

NO

OH•54-70%

533:1 dr

NO

OR•

TMS5450

Scheme 11. Propargylation of aldehyde 50

In reexamining the propargylation step, it seemed necessary to install the propargyl group

and to protect the terminal alkyne in a single step. There are relatively few methods of

accomplishing this. Most propargylation conditions rely upon the favorable equilibrium of the

propargylic and allenic organometallic species, 56 and 57, respectively (Scheme 12), the latter

leading to the α-addition product 58 via an SE2’ mechanism.85 Sterically demanding groups on

the terminus of the alkyne (e.g., R1 = TMS) may shift this equilibrium to favor the propargyl

organometallic species 56, giving the allenic alcohol 55 preferentially. The product distribution

has, however, been found to be highly dependent on the metal used, the nature of the substituent

on the alkyne (R1), the electrophile, and reaction solvent.

22

R1M

•R1

M

α β

γ R2CHO

R2

OH R1

R2

OH•

R1

R2CHO

55 56 57 58

α βγ

Scheme 12. Regioselectivity in propargylation reactions

Paquette and Daniels86 have published an often-cited study that exemplifies the

capricious behavior of propargylation reactions. Treatment of aldehydes and ketones with

trimethylsilylpropargylzinc bromide in THF led to selective formation of the homopropargylic

alcohol, while treatment of the same substrates with the analogous aluminum reagent yielded

allenyl alcohols preferentially. Additionally, increasing the coordinating ability of the solvent

was highly influential in the product distribution of protonolysis of the aluminum reagent; in

diethyl ether, 1-trimethylsilylpropyne was obtained after quenching the organometallic species,

while in THF or diglyme the regioselectivity was reversed.

Unfortunately, Paquette and Daniels’87 zinc amalgam protocol failed to produce the

desired homopropargyl alcohol 54 (Table 3, entry 1). The majority of the starting material 50

was recovered. Propargylation of aldehyde 50 was attempted using conditions developed by Loh

and Lin,88 who have shown that simple aldehydes undergo regioselective propargylation with

trialkylsilyl propargyl bromides in the presence of indium and catalytic indium trifluoride. Lin

and Loh attribute the high selectivity for the α-alkylation product to the probable coordination of

the halogen and the silicon in the allenic organometallic species 57. This protocol proved

ineffective in the propargylation of our aldehyde 50, however (entry 2). The majority of the

starting material was recovered.

23

Table 3. Propargylation of aldehyde 50

NO

O•

Br

conditions

NO

OR2•

5450TMS

R1

Entry R1 Conditions R2 Yield of 54 (%)

1 TMS Zn, HgCl2, cat I2 NA NR

2 TMS In, InF3, THF, reflux NA NR

3 H 2 equiv n-BuLi, TMEDA, then TMSCl TMS 21 (2:1 dr)

Cabezas89, 90 has also presented an approach to the problem of regioselective addition, in

which 1,3-dilithiopropyne is prepared from propargyl bromide and added to the aldehyde.

Quenching of the reaction with TMSCl protects both the newly formed alcohol and the alkyne

terminus. Employment of this method gave the desired bis-protected homopropargyl alcohol 54

(R2 = TMS), albeit in 21% yield as a 2:1 mixture of diastereomers by 1H NMR (Table 3, entry

3).

Gratifyingly, when subjected to [Rh(CO)2Cl]2 in the presence of CO, allene-yne 54

smoothly underwent cyclocarbonylation to give the desired product 59 in 91% yield after 3 hours

(Scheme 13). Following cyclocarbonylation, the two diastereomers (originating in the

propargylation step) were easily separated by column chromatography, yielding 10 mg (73%) of

the major diastereomer, and 3 mg (18%) of the minor isomer.

24

NO

OTMS

54

TMS

OTMS

O

NO

30 mol% [Rh(CO)2Cl]2

CO, toluene, 60 °C91%

TMS59

Scheme 13. Cyclocarbonylation of the model system

There is a minor discrepancy between the isolated yields of the diastereomers of 59 and

the previously observed 2:1 dr observed in the 1H NMR of the homopropargyl silyl ether 54.

Were the remaining 9% of the material not isolated as product comprised solely of the minor

diastereomer of 54, the diastereomeric ratio observed in the cyclocarbonylation reaction

(potentially 73:27) still would not be equivalent to 2:1. This lack of agreement may be attributed

to two things. First, the 2:1 dr is based on the integration of methine protons that are not fully

resolved in the 1H NMR spectrum (see Appendix A). Second, the margin of error in weighing

milligram quantities is presumably large enough to account for the observed discrepancy.

In an effort to gain insight into the relative stereochemistry of the cyclocarbonylation

products, the two epimers of 59 were modeled using CAChe. The global minimum

conformation for each isomer was found using CONFLEX/MM3 parameters (Figure 6).91 The

minimum energy conformer for the epimer with the α-TMS ether (59α) exhibited dihedral

angles of 70.2 and 44.3° between Ha and Hb, and Ha and Hc, respectively. The measured dihedral

angles for 59β are 176.2 and 60.9°. Based on these calculations, one would expect to observe a

doublet of doublets with two medium-to-small 3J couplings for the resonance of Ha in 59α. In

contrast, the calculated dihedral angles for 59β predict a doublet of doublets with one large and

one small vicinal coupling. Indeed, the resonance for Ha in the 1H NMR spectrum of the major

diastereomer of 59 is a triplet at δ 5.01 ppm with J = 3.6 Hz (see Appendix A). The

25

corresponding signal in the 1H NMR spectrum of the minor diastereomer is a doublet of doublets

at δ 4.76 ppm with J = 10.8 and 5.4 Hz. Therefore, the major and minor diastereomers can

tentatively be assigned as 59α and 59β, respectively.

O

NO

TMS59α

OTMSHa

HcHb O

NO

TMS59β

Ha

OTMS

HcHb

Ha-C-C-Hb ∠ 70.2°Ha-C-C-Hc ∠ 44.3°

observed for Ha: δ 5.01 (t, J = 3.6 Hz)

Ha-C-C-Hb ∠ 176.2°Ha-C-C-Hc ∠ 60.9°

observed for Ha: δ 4.75 (dd, J = 10.8, 5.4 Hz)

Figure 6. Minimum energy conformations and dihedral angles for 59α (left) and 59β (right)

4.3 SYNTHESIS OF AN ASYMMETRICALLY-FUNCTIONALIZED C RING

Our strategy for the synthesis of a functionalized C ring was to add isopropenyl Grignard to

known enone 60 (Scheme 14), hydrate the exocyclic olefin of 62 and cleave the ketal to access

functionalized enone 28.

26

O

MeO OMe MeO OMe

OH OHPO

O60 61 28

Scheme 14. Proposed synthesis of functionalized C ring

4.3.1 Conjugate addition to cyclohexadienone 62 to form enone 60

There are only a few literature examples of conjugate addition of a methyl group to quinone

monoketal 62. Addition of lithium dimethylcopper to quinone monoketal 6292 is known to lead

to reductive aromatization rather than conjugate addition (Table 4, entry 1).93 Complexation of

quinone monoketals with an organoaluminum reagent followed by addition of either an

alkyllithium or Grignard reagent has been shown to yield the corresponding 1,4-addition

products in many cases; however, in the case of the addition of methyllithium, only a 24% yield

of the desired product 60 was achieved (entry 2).94 Fortunately, Feringa and colleagues95 have

developed an enantioselective copper-catalyzed conjugate addition protocol that was reported to

give the desired enone 60 in 76% yield and 99% ee from quinone monoketal 62 (entry 3).

27

Table 4. Literature examples of conjugate addition to quinone monoketal 62

O

MeO OMe62

conditionsO

MeO OMe60

Entry Conditions Yield (%) Ref.

1 LiMe2Cu —a 93

2 MAD,b MeLi 24 94

3 Cu(OTf)2, (S,R,R)-64, Me2Zn 76 (99% ee) 95 ap-Methoxyphenol obtained exclusively; bMAD =

methylaluminum bis(2,6)-di-tert-butyl-4-methylphenoxide)

The enantioselectivity of Feringa’s protocol arises from the use of chiral phosphoramidite

ligand 64 (Figure 7). The Feringa group has published the use of a variety of phosphoramidite

ligands synthesized from BINOL, PCl3, and 2° amines.96 Because the price of the amines varies

widely, we sought to synthesize an effective phosphoramidite ligand while minimizing cost. We

were encouraged to see that both ligand (S)-63 and (S,R,R)-64 were effective in the copper-

catalyzed addition of diethylzinc to 2-cyclohexenone, each giving 3-ethylcyclohexanone in

>75% yield and 83% ee and >98% ee, respectively.96 Initially less concerned with the

enantiopurity of the enone than with obtaining the desired material, we opted to synthesize

racemic phosphoramidite 63 from racemic BINOL (rac-65), PCl3, and diisopropylamine

(Scheme 15).

OP

ON

OP

ON

Ph

Ph

(S,R,R)-64(S)-63

Figure 7. Chiral phosphoramidite ligands

28

OHOH

rac-65

PCl3, TEA

toluene, -60 °C OP

OCl

rac-66

DIPA, TEA

toluene, -40 °C → rt19% O

PO

N

rac-63

Scheme 15. Synthesis of racemic phosphoramidite ligand 63

Unfortunately, exposure of quinone monoketal 62 to Cu(OTf)2, ligand rac-63, and Me2Zn

in toluene at –25 °C failed to produce the desired product 60 in a reasonable yield (Table 5,

entries 1 and 2). The reaction did not go to completion, even when the mixture was allowed to

stir at -25 °C for several days. The lack of desired reactivity may have been due, at least in part,

to the poor solubility of the catalyst complex in toluene at the reaction temperature. Upon

cooling the premixed solution of Cu(OTf)2 and ligand from room temperature to -25 °C,

significant precipitation was observed.

Enantiomerically pure phosphoramidite (S,R,R)-64 was synthesized in 26% yield in a

fashion analogous to the previous phosphoramidite synthesis.96 Cu(OTf)2 was premixed with the

phosphoramidite ligand and was then cooled to -25 °C; the quinone monoketal was added

followed by a solution of Me2Zn. Within approximately six hours, all cyclohexadienone 62

starting material was consumed (this is in contrast to the 16 hours the reaction is reputed95 to

require). Surprisingly, a significant amount of the diaddition product 67 was observed in the

product mixture (Table 5, entry 3). Protonation of the Zn-enolate intermediate prior to

quenching of the catalyst is requisite in order for a second 1,4-addition to occur. TLC indicated

a dramatic difference in the relative amounts of mono- and diaddition products (60 and 67,

respectively) before and after quenching. The final ratio of mono- to diaddition products has

been found to be somewhat contingent on the method of quenching. This trend is summarized in

Table 5 (entries 3-5), although it is important to note that exact ratios were not reproducible.

29

Following the published procedure,95 the addition of saturated aqueous NH4Cl to the reaction

mixture led to greater relative amounts of the diaddition product 67 in the crude 1H NMR than

did a basic quench (entries 3 and 4), which was prescribed by Rosalinde Imbos in her Ph.D.

thesis.97 The ratio of mono- to diaddition product (60 : 67) was further improved by performing a

reverse basic quench (entry 5).

Table 5. Product distribution in copper-catalyzed conjugate addition to cyclohexadienone 62

O

MeO OMe

2.5 mol% Cu(OTf)25 mol% ligand

toluene, -25 °C

62

O

MeO OMe60

O

MeO OMe67

*+ + unreacted

62

Entry Ligand Equiv Me2Zn

Time (h) Quench 60 : 67 : 62a Yield of 60

(%)

1 rac-63 1.6 16 direct, sat NH4Cl — 17

2 rac-63 1.6 + 1b 48 direct, sat NH4Cl — 20

3 (S,R,R)-64 1.6 6 direct, sat NH4Cl 18 : 82 : 0 —c

4 (S,R,R)-64 1.6 16 direct, 1 M NaOH 60 : 40 : 0 33

5 (S,R,R)-64 1.6 6 reverse, 1 M NaOH 71 : 29 : 0 —c

6 (S,R,R)-64 1.6 4d reverse, 1.5 M NaOH 75 : 0 : 25 50

7 (S,R,R)-64e 1.6 16 direct, 1 M NaOH 100 : 0 : 0 21f adetermined by integration of crude 1H NMR; badditional equivalent Me2Zn added after 45 h; cyield not determined; dreaction followed closely by TLC and quenched at first faint sign of diaddition product; eligand of high purity; flow yield due to volatility of product

The presence of the diaddition product proved to be problematic since separation of the

mono- and diaddition products by column chromatography was not possible. Two methods of

minimizing the formation of the diaddition product were discovered. The first was to monitor

the reaction closely and quench the reaction as soon as there was TLC evidence of diaddition

product forming (Table 5, entry 6). An early quench of the reaction allowed for the isolation of

an easily separable 3:1 mixture (by 1H NMR) of desired product 60 to unreacted starting

30

material. (Fortuitously, ketone 67 stains very darkly with PAA and although the diaddition

product was evident by TLC, it often was only faintly—if at all—visible in the crude 1H NMR.)

The use of phosphoramidite ligand of a higher purity also suppressed the formation of the

unwanted diaddition product.98 Unsatisfied with the previously low-yielding synthesis of

phosphoramidite 64, we developed a revised synthesis based on methods reported in Ate

Duursma’s Ph.D. thesis.99 (S)-BINOL (65) was refluxed in PCl3 (Scheme 16). Removal of the

excess PCl3 under vacuum furnished phosphochloridite 66, which was dissolved in toluene and

added to a solution of deprotonated amine (R,R)-68. Not only did this method drastically

improve the isolated yield of the ligand 64 (77% versus 26%), but also the ligand was

significantly more pure by 1H NMR.100 Use of this ligand in the copper-catalyzed conjugate

addition led only to the desired monoaddition product (Table 5, entry 7). Unfortunately,

repetition of the improved ligand synthesis procedure led to ligand of inferior purity that, in turn,

led to significant quantities of diaddition product 67 produced in the copper-catalyzed 1,4-

addition reaction.

OHOH

(S)-65

reflux OP

OCl

(S)-66

toluene/THF, -40 °C → rt77% O

PO

N

(S,R,R)-64

Ph

Ph

PCl3

NPh

PhLi

(R,R)-68

Scheme 16. Improved synthesis of phosphoramidite (S,R,R)-64

31

4.3.2 Elaboration of enone 60

Isopropenylmagnesium bromide was added to the ketone 60, affording the tertiary alcohol 61 in

74% yield (Scheme 17). Hydroboration of the 1,1-disubstituted olefin with BH3·DMS followed

by oxidation101 gave the corresponding primary alcohol as a mixture of diastereomers. Attempts

to improve the diastereoselectivity by using bulky 9-BBN led to no appreciable hydroboration, as

observed by the 1H NMR spectra of aliquots of the reaction mixture. The mixture of

diastereomers was taken on crude due to the lability of the ketal on silica gel, even when the

column was pretreated with TEA. The primary alcohol was protected as the TBS ether102 prior

to cleavage of the ketal by Montmorillonite K10 clay,103 giving the enone 28 in 21% yield from

61 as a 2:1 mixture of diastereomers by 1H NMR.

O

MeO OMe60

*

MgBr

THFMeO OMe

61

OH 1. BH3·DMS, then H2O, NaOH, H2O2

28

OH

2. TBSCl, DMAP, TEA3. Montmorillonite-K10

TBSO

O21%, 3 steps74%

Scheme 17. Synthesis of enone 28

A variety of 1,3-dipolar cycloaddition conditions were surveyed, all of which were

designed to generate the nitrile oxide very gradually to minimize homodimerization of the nitrile

oxide.56 Unfortunately, enone 28 failed to undergo 1,3-dipolar cycloaddition and, in all cases,

most of the starting material was recovered (Table 6). A very small amount of a mixture of

unknown products was isolated from the reaction of enone 28 with nitroalkane 43 when Kurth’s

diisocyanate methodology75 was used (entry 6). Since selectivity was not exhibited,

investigation was discontinued.

32

Table 6. Attempted stereoselective cycloaddition of enone 26

OHTBSO

O

OHTBSO

O

NO

R28 26

Ph Cl

NOH

NO2TBSO

29 43

or

conditions

Entry 29 or 43 Conditions Scale (mg of 28)

Yield of 26 (%)

1 29 EtMgBr, DCM70 10 —

2 29 TEA, Et2O71 10 —

3 29 KHCO3, DME104 10 —

4 29 KHCO3, DME, Δ 10 —

5 29 4 Å mol sieves, DCM105 21 —

6 43 1,4-phenylene diisocyanate, TEA, PhH, reflux75 21 —a aMost of starting material was unreacted; however, a small amount of a mixture of unknown products was obtained.

33

5.0 CONCLUSION

Our proposed plan for the synthesis of resiniferatoxin required that we addressed several

uncertainties, namely whether an isoxazoline moiety would be tolerated by the Rh(I)-catalyzed

cyclocarbonylation reaction, whether selective oxidations of the resulting alkylidene

cyclopentenone could be carried out, and whether a stereoselective 1,3-dipolar cycloaddition was

feasible. We synthesized a simple isoxazoline-containing model system, which underwent

Rh(I)-catalyzed allenic cyclocarbonylation to give the unfunctionalized angular [6-7-5] core of

resiniferatoxin in 91% yield. The cyclopentenone scaffold produced by the cyclocarbonylation

reaction was an appropriate substrate for probing conditions for selective olefin oxidation,

although selective oxidation was not pursued due to the disappointing results obtained in the area

of stereoselective 1,3-dipolar cycloaddition. An asymmetrically functionalized enone,

corresponding to the C ring of resiniferatoxin, was synthesized utilizing Feringa’s Cu-catalyzed

asymmetric conjugate addition protocol. Unfortunately, the functionalized 2-cyclohexenone

failed to undergo stereoselective 1,3-dipolar cycloaddition under numerous conditions tested.

34

6.0 EXPERIMENTAL

6.1 GENERAL

Unless otherwise specified, all nonaqueous reactions were performed in flame- or oven-dried

glassware under N2 atmosphere using the appropriate syringe, cannula, and septum techniques.

All solvents and reagents were purchased commercially and used as received unless otherwise

noted. Toluene, acetonitrile (MeCN), triethylamine (TEA), N,N,N’,N’-

tetramethylethylendiamine (TMEDA), and chlorotrimethylsilane (TMSCl) were distilled from

CaH2 prior to use. Dimethyl sulfoxide (DMSO) was distilled from CaH2 and stored over

activated 4 Å mol sieves. Tetrahydrofuran (THF) and diethyl ether (Et2O) were dried and

deoxygenated by sequential passage through activated alumina and Q5 columns in a Sol-Tek ST-

002 solvent purification system; dichloromethane (DCM) was passed through an activated

alumina column in the same system. 2-Cyclohexenone was distilled (47 °C at 9 mm Hg) and

stored at -20 °C. Acetic anhydride (Ac2O) was distilled (30 °C at 20 mm Hg) from P2O5 and

stored in a desiccator. Cu(OTf)2 and [(Ph3P)CuH]6 were stored in a glove box and dispensed

using standard glove box techniques. All solvents used in the preparation and reactions of air-

sensitive reagents were degassed by bubbling N2 for 20-30 min before use.

Column chromatography was performed using silica gel (32-63 μm particle size, 60 Å

pore size) purchased from Scientific Adsorbents, Inc. following the guidelines described in the

35

seminal publication on flash chromatography by Still and colleagues.106 Ethyl acetate (EtOAc)

and hexanes used for chromatography were distilled prior to use. TLC was performed using

silica gel plates (60 F254, 250 μm thickness).

1H and 13C NMR spectra were obtained on Varian 300 MHz instruments. All chemical

shifts (δ) are reported in ppm. 1H NMR spectra were calibrated to the residual CHCl3 peak at δ

7.27; 13C NMR spectra were referenced to the CDCl3 resonance at δ 77.0. The following

abbreviations are used to denote the indicated splitting pattern 1H NMR spectra: s = singlet, d =

doublet, t = triplet, q = quartet; abbreviations are used in combination to indicate more complex

splitting (e.g., dtd = doublet of triplets of doublets). Infrared spectra were obtained on a Nicolet

Avatar E. S. P. 360 FT-IR.

6.2 SYNTHESIS AND CHARACTERIZATION

OTBSO2N

43

OCN

NCO

TEA, PhH, refluxO

NO

OTBS45

2-cyclohexenone (44)

3-((tert-Butyldimethylsilyloxy)methyl)-5,6,7,7a-tetrahydrobenzo[d]isoxazol-4(3aH)-one

(45). A mixture of cyclohexenone (44) (1.80 mL, 18.59 mmol, 1.0 equiv), nitroalkane 43 (3.75

g, 18.28 mmol, 1.0 equiv), 1,4-phenylenediisocyanate (9.08 g, 56.69 mmol, 3.1 equiv) and TEA

(0.50 mL, 3.59 mmol, 0.2 equiv) in benzene (180 mL) was refluxed. After approximately 16 h,

the turbid yellow mixture was cooled to rt; water (9 mL) was added and the mixture stirred 1 h.

The polymeric urea was removed via vacuum filtration through a pad of celite; the filtrate was

36

dried (MgSO4) and concentrated under reduced pressure. Benzene was added and the filtration

step was repeated. The filtrate was concentrated onto silica gel, which was dry-loaded onto a

silica gel column for purification. Gradient elution with EtOAc/hexanes (1:9 to 1:6) yielded

isoxazoline 45 (2.212 g, 43%) as a colorless oil. 1H NMR (300 MHz, CDCl3): δ 4.95 (dt, J =

9.7, 4.5 Hz, 1H), 4.56 (A of ABq, J = 12.5 Hz, 1H), 4.44 (B of ABq, J = 12.5 Hz, 1H), 3.86 (d, J

= 9.7 Hz, 1H), 2.54-2.45 (m, 1H), 2.37-2.26 (m, 1H), 2.16-1.79 (m, 4H), 0.89 (s, 9H), 0.10 (s,

3H), 0.09 (s, 3H); 13C NMR (75 MHz, CDCl3): δ 205.3, 157.4, 81.5, 58.7, 58.5, 39.9, 26.4, 25.8,

18.4, 18.3, -5.5, -5.6; IR (neat, NaCl): 2930, 2857, 1713, 1088, 837 cm-1; MS m/z (%): 283 (9),

268 (73), 226 (90), 74 (100); EI-HRMS calcd for C14H25NO3Si [M]+ m/z: 283.1604, found:

283.1610.

NO

OTBS

46

THF, 0 °C

MgBr

HO84%

O

NO

OTBS

45

3-((tert-Butyldimethylsilyloxy)methyl)-4-ethynyl-3a,4,5,6,7,7a-hexahydrobenzo[d]isoxazol-

4-ol (46). To a solution of the ketone 45 (0.511 g, 1.802 mmol, 1 equiv) in THF (18 mL) at 0 °C

was added ethynylmagnesium bromide (14 mL of a 0.5 M soln in THF, 7 mmol, 4 equiv) over

25 min via syringe pump. The mixture stirred at 0 °C and reaction progress was monitored by

TLC. Complete consumption of the starting material was observed ~30 min after the addition

was finished; the reaction was quenched by pouring the mixture into a flask containing sat

NH4Cl solution. The aqueous layer was extracted with Et2O (4×) and the combined organic

layers were washed with brine, dried (MgSO4), filtered, and concentrated under reduced

pressure. The residue was purified via column chromatography, using EtOAc/hexanes (1:6) as

the eluent, to give pure propargyl alcohol 46 (0.470 g, 84%) as a colorless oil. 1H NMR (300

37

MHz, CDCl3): δ 4.67-4.61 (m, 1H), 4.62 (A of ABq, J = 11.7 Hz, 1H), 4.45 (B of ABq, J = 11.7

Hz, 1H), 3.53 (d, J = 9.0 Hz, 1H), 2.48 (s, 1H), 1.97-1.71 (m, 4H), 1.61-1.50 (m, 1H), 0.93 (s,

9H), 0.16 (s, 6H); 13C NMR (75 MHz, CDCl3): δ 157.6, 87.2, 80.2, 71.6, 65.4, 58.6, 57.9, 36.6,

25.8, 25.0, 18.2, 16.1, -5.5; IR (neat, NaCl): 3416, 3310, 2930, 2090, 1257, 1081, 838 cm-1; MS

m/z (%): 309 (9), 294 (69), 252 (82), 105 (74), 74 (100); EI-HRMS calcd for C16H27NO3Si [M]+

m/z: 309.1760, found: 309.1762.

NO

OTBS

46

HO DCM, 0 °C

NO

OTBS

47

AcO

Ac2O, TEA, DMAP

74%

3-((tert-Butyldimethylsilyloxy)methyl)-4-ethynyl-3a,4,5,6,7,7a-hexahydrobenzo[d]-isoxazol-

4-yl acetate (47). To a soln of the propargyl alcohol 46 (0.560 g, 1.81 mmol, 1 equiv) and

DMAP (0.230 g, 1.88 mmol, 1 equiv) in DCM (18 mL) at 0 °C were added Ac2O (1.7 mL, 18.00

mmol, 10 equiv) and TEA (2.5 mL, 17.94 mmol, 10 equiv). The mixture stirred at 0 °C and

reaction progress was monitored by TLC. After approximately 12 h, the reaction was quenched

by addition of saturated NaHCO3 soln and the aqueous layer was extracted with DCM (3×). The

combined organic layers were washed with brine, dried (MgSO4), filtered, and concentrated

under reduced pressure. The residue was purified by column chromatography, eluting with

EtOAc/hexanes (1:6), to give pure propargyl acetate 47 (0.504 g, 74%) as a colorless oil. 1H

NMR (300 MHz, CDCl3): δ 4.58 (A of ABq, J = 12.7 Hz, 1H), 4.52 (B of ABq, J = 12.7 Hz,

1H), 4.54-4.50 (m, 1H), 3.58 (d, J = 8.4 Hz, 1H), 2.72-2.66 (m, 1H), 2.68 (s, 1H), 2.12-2.04 (m,

1H), 2.03 (s, 3H), 1.87-1.57 (m, 5H), 0.91 (s, 9H), 0.11 (s, 3H), 0.10 (s, 3H); 13C NMR (75 MHz,

CDCl3): δ 168.7, 159.8, 82.8, 79.0, 74.9, 70.8, 58.7, 54.3, 32.4, 25.8, 24.2, 21.8, 18.2, 15.3, -5.2,

-5.4; IR (neat, NaCl): 3250, 2931, 2100, 1748, 1221, 1076, 838 cm-1; MS m/z (%) 294 (85), 234

38

(100), 156 (20), 117 (41), 105 (24); EI-HRMS calcd for C17H26NO4Si [M]+ m/z: 336.1631,

found: 336.1631.

NO

OTBS

47

AcO toluene, H2O

[(Ph3P)CuH]6

87%

NO

OTBS

48

3-((tert-Butyldimethylsilyloxy)methyl)-4-vinylidene-3a,4,5,6,7,7a-

hexahydrobenzo[d]isoxazole (48). To a dark red suspension of [(Ph3P)CuH]682 (1.68 g, 0.86

mmol, 0.50 equiv) in toluene (8.5 mL) was added a mixture of the propargyl acetate 47 (0.607 g,

1.727 mmol, 1 equiv), toluene (17 mL) and water (0.08 mL, 4.44 mmol, 2.57 equiv) via cannula

over 20 min. The mixture gradually turned brown (i.e., appeared quenched) approximately 1 h

after the addition. The septum was removed from the flask and the contents stirred open to the

air for 20 min before being diluted with Et2O and filtered through a plug of silica gel. After

concentration under reduced pressure, the filtrate was purified by column chromatography using

gradient elution with hexanes/EtOAc (19:1, 9:1). A second purification using 19:1

hexanes/EtOAc was required to remove remaining PPh3. Allene 48 (0.305 g, 87%) was obtained

as a colorless oil. 1H NMR (300 MHz, CDCl3): δ 4.72-4.60 (m, 3H), 4.46 (A of ABq, J = 13.2

Hz, 1H), 4.42 (B of ABq, J = 13.2 Hz, 1H), 3.86 (br d, J = 9.3 Hz, 1H), 2.17-2.12 (m, 2H), 1.79-

1.62 (m, 4H), 1.53-1.45 (m, 1H), 0.87 (s, 9H), 0.07 (s, 3H), 0.06 (s, 3H); 13C NMR (75 MHz,

CDCl3): δ 205.6, 160.2, 95.3, 79.7, 75.2, 58.0, 49.0, 27.2, 26.1, 25.7, 19.3, 18.1, -5.5, -5.6; IR

(neat, NaCl): 2953, 2830, 1960, 1255, 1088, 839 cm-1; MS m/z (%): 278 (41), 239 (33), 238 (96),

237 (73), 236 (100), 156 (13); EI-HRMS calcd for C15H24NO2Si [M]+ m/z: 278.1576, found:

278.1576.

39

NO

OTBS

48

•THF, 0 °C

TBAF NO

OH

49

•89%

(4-Vinylidene-3a,4,5,6,7,7a-hexahydrobenzo[d]isoxazol-3-yl)methanol (49). To a soln of the

TBS ether 48 (0.166 g, 0.565 mmol, 1 equiv) in THF (5.7 mL) at 0 °C was added TBAF (0.85

mL of a 1.0 M soln in THF, 0.85 mmol, 1.5 equiv) dropwise. Mixture stirred at 0 °C and the

reaction progress was monitored by TLC. Approximately 20 min after the addition was

complete, the reaction was quenched by the addition of sat. NH4Cl soln. The aqueous layer was

extracted with Et2O (4×). The combined organic layers were washed with brine, dried (MgSO4),

filtered, and concentrated under reduced pressure. The crude residue was purified via column

chromatography, using EtOAc/hexanes (3:7), to give alcohol 49 (0.075 g, 75%) as a pale yellow

oil. 1H NMR (300 MHz, CDCl3): δ 4.72-4.71 (m, 2H), 4.64 (dt, J = 9.1, 5.1 Hz, 1H), 4.42 (A of

ABq, J = 14.7 Hz, 1H), 4.34 (B of ABq, J = 14.7 Hz, 1H), 3.84 (br d, J = 9.1 Hz, 1H), 3.25 (br s,

1H), 2.20-2.10 (m, 2H), 1.78-1.58 (m, 3H), 1.53-1.40 (m, 1H); 13C NMR (75 MHz, CDCl3): δ

205.4, 161.0, 95.0, 79.7, 75.6, 57.6, 49.1, 27.1, 26.1, 19.4; IR (neat, NaCl): 3392, 2941, 1959,

1610, 1442, 1037 cm-1.

NO

OH

49

•TEA, DCM, -65 °C

(COCl)2, DMSO NO

O

50

•75%

4-Vinylidene-3a,4,5,6,7,7a-hexahydrobenzo[d]isoxazole-3-carbaldehyde (50). To a soln of

oxalyl chloride (0.08 mL, 0.932 mmol, 2.1 equiv) in DCM (3 mL) at –65 °C was added a soln of

DMSO (0.14 mL, 1.971 mmol, 4.4 equiv) in DCM (1.5 mL) dropwise over 10 min. The mixture

stirred 10 min before a soln of the alcohol 49 (0.080 g, 0.446 mmol, 1 equiv) in DCM (1.5 mL)

40

was added dropwise over 10 min. The mixture became cloudy while stirring for approximately

20 min at –65 °C. TEA was added dropwise over 5 min and the mixture was allowed to warm to

room temperature over 30 min. The reaction mixture was washed with sat NH4Cl soln and

water. The combined aqueous washings were back-extracted with DCM (3×). The combined

organic layers were washed with brine, dried (MgSO4), filtered, and concentrated under reduced

pressure. The crude residue was purified via column chromatography, eluting with

hexanes/EtOAc (6:1), to give aldehyde 50 (0.059 g, 75%) as a colorless oil. 1H NMR (300 MHz,

CDCl3): δ 9.90 (s, 1H), 4.75-4.64 (m, 2H), 3.94 (dt, J = 8.7, 3.6 Hz, 1H), 2.30-2.01 (m, 3H),

1.88-1.77 (m, 1H), 1.71-1.60 (m, 2H); 13C NMR (75 MHz, CDCl3): δ 204.9, 185.4, 162.2, 95.4,

84.5, 76.9, 44.4, 26.9, 24.7, 19.4; IR (neat, NaCl): 2946, 2843, 1960, 1696, 1160 cm-1; MS m/z

(%): 177 (65), 160 (82), 106 (87), 105 (100), 104 (70); EI-HRMS calcd for C10H11NO2 [M]+ m/z:

177.0790, found: 177.0781.

NO

O•

Br

Zn, sat NH4Cl, THF

NO

OH•54-70%

5350

1-(4-Vinylidene-3a,4,5,6,7,7a-hexahydrobenzo[d]isoxazol-3-yl)but-3-yn-1-ol (53). To a slurry

of Zn dust (0.046 g, 0.701 mmol, 5 equiv) in THF (0.2 mL) at 0 °C, a soln of the aldehyde 50

(0.025 g, 0.140 mmol, 1 equiv) in THF (0.2 mL) and propargyl bromide (0.04 mL of an 80 wt%

soln in toluene, 0.359 mmol, 2.5 equiv) were added, followed by the addition of sat NH4Cl soln

(0.16 mL) dropwise over 45 min. The reaction mixture was allowed to slowly warm to room

temperature and stirred overnight. After approximately 20 hours, TLC indicated complete

consumption of starting material and the reaction mixture was filtered through a pad of Celite.

41

The filtrate was washed with sat NH4Cl soln and brine, dried (Na2SO4), filtered, and

concentrated under reduced pressure. The residue was purified via column chromatography

using hexanes/EtOAc (3:1) as the eluent. The homopropargyl alcohol 53 (0.021 g, 70%) was

achieved as a pale yellow crystalline solid and an approximate 3:1 mixture of diastereomers (1H

NMR). 1H NMR (300 MHz, CDCl3): δ 4.78-4.75 (m, 3H), 4.68-4.66 (m, 1H), 4.00 (br d, J =

9.3, 0.2 H),* 3.89 (br d, J = 9.0, 0.8 H),** 2.96 (br d, J = 5.7, 1H), 2.75 (A of ABqdd, J = 17.1,

5.4, 2.7, 1H), 2.65 (B of ABqdd, J = 17.1, 5.7, 2.7, 1H), 2.26-2.13 (m, 2H), 2.08 (t, J = 2.7, 1H),

1.86-1.48 (m, 4H); 13C NMR (75 MHz, CDCl3): δ 205.8, 161.1, 95.3, 80.6, 78.9, 75.8, 71.6,

66.8, 49.8, 27.2, 26.3, 25.6, 19.5; IR (neat, NaCl): 3390, 3292, 2946, 2120, 1958, 1059. (*Minor

diastereomer; **major diastereomer.)

NO

O•

Br

2 equiv n-BuLi, TMEDA, then TMSCl21%, 2:1 dr

NO

OTMS•

5450TMS

3-(4-(Trimethylsilyl)-1-(trimethylsilyloxy)but-3-ynyl)-4-vinylidene-3a,4,5,6,7,7a-

hexahydrobenzo[d]isoxazole (54). n-BuLi (0.5 mL of a 1.6 M soln in hexanes, 0.8 mmol, 4.2

equiv) was added to a mixture of Et2O (0.7 mL) and hexanes (0.3 mL) at -78 °C. TMEDA (30

μL, 0.2 mmol, 1.05 equiv) was added dropwise followed by slow dropwise addition of propargyl

bromide (45 μL of an 80 wt% soln in toluene, 0.5 mmol, 2.1 equiv). A white precipitate

developed upon stirring at –78 °C 20 min. A soln of the aldehyde 50 (0.034 g, 0.192 mmol, 1

equiv) in Et2O (0.25 mL) was added dropwise and the mixture continued to stir at –78 °C until

complete consumption of starting material was observed by TLC. The reaction was then

quenched by rapid addition of TMSCl (0.10 mL, 0.79 mmol, 4 equiv). The mixture was allowed

42

to slowly approach room temperature until no additional progress was observed by TLC, at

which point the mixture was poured into an Erlenmeyer flask containing ~10 mL H2O. The

aqueous layer was extracted with Et2O (4×). The combined organic portions were dried

(Na2SO4), filtered, and concentrated under reduced pressure. The residue was purified by

column chromatography, using hexanes/EtOAc (9:1) as the eluent, to give allene-yne 54 (0.014

g, 21%) as a 2:1 mixture of diastereomers (1H NMR). 1H NMR (300 MHz, CDCl3): δ 4.74-4.67

(m, 4H), 3.98 (d, 9.9 Hz, 0.3H),* 3.92 (d, J = 9.6 Hz, 0.7H),** 2.78 (A of ABqd, J = 16.8, 5.7

Hz, 1H), 2.65 (B of ABqd, J = 16.8, 7.8 Hz, 1H), 2.28-2.20 (m, 2H), 1.80-1.35 (m, 4H), 0.20 (s,

9H), 0.15 (s, 9H); 13C NMR (75 MHz, CDCl3): δ 206.1, 160.2, 103.2, 95.8, 86.8, 80.2, 75.3,

67.4, 49.6, 27.0, 26.5, 26.4, 19.3, 0.3, 0.0; IR (neat, NaCl): cm-1; TOF-HRMS calcd for

C19H31NO2Si2 [M+H]+ m/z: 362.1972, found: 362.1955. (*Minor diastereomer; **major

diastereomer.)

NO

OTMS

54

TMS

OTMS

O

NO

30 mol% [Rh(CO)2Cl]2

CO, toluene, 60 °C91%

TMS59

Cyclopentenone 59. To a test tube containing the allene-yne 54 (0.013 g, 0.036 mmol) and a stir

bar was added toluene (0.4 mL). The test tube was evacuated and flushed with CO (3×).

[Rh(CO)2Cl]2 (0.005 g, 0.012 mmol, 0.33 equiv) was added in one portion and the test tube was

evacuated and flushed with CO (3×) before being placed under CO (1 atm) and immersed in a 60

°C bath. TLC analysis indicated near complete consumption of starting material after 3.5 h, at

which point the reaction mixture was flushed through a plug of silica gel with hexanes/EtOAc

(1:1) to remove the Rh catalyst. The collected eluent was concentrated under reduced pressure

43

and the residue was purified via column chromatography using hexanes/EtOAc (3:1) as the

eluent, giving the major diastereomer of cyclopentenone 59 (0.010 g, 73%) followed by the

minor diastereomer (0.003 g, 21%). Both were white crystalline solids. Major diastereomer: 1H

NMR (300 MHz, CDCl3): δ 5.01 (t, J = 3.6 Hz, 1H), 4.55 (dt, J = 8.2, 3.0 Hz, 1H), 3.81 (br d, J

= 8.2 Hz, 1H), 3.39 (dd, J = 14.4, 3.6 Hz, 1H), 2.95 (A of ABq, J = 20.4 Hz, 1H), 2.88 (B of

ABq, J = 20.4, 1H), 2.85 (dd, J = 14.4, 3.6 Hz, 1H), 2.58-2.53 (m, 1H), 2.28-2.24 (m, 1H), 1.87-

1.67 (m, 4H), 0.28 (s, 9H), 0.20 (s, 9H); 13C NMR (75 MHz, CDCl3): δ 208.2, 173.4, 162.7,

135.8, 132.8, 80.4, 66.3, 49.3, 41.6, 37.3, 31.0, 25.1, 19.2, 0.2, 0.1; IR (neat, NaCl): 2950, 2880,

1676, 1549, 1246, 839 cm-1; MS m/z (%): 389 (11), 374 (61), 147 (29), 75 (74), 73 (100); HR-

EIMS calcd for C20H31NO3Si [M]+ m/z: 389.1843, found: 389.1839. Minor diastereomer: 1H

NMR (300 MHz, CDCl3): δ 4.76 (dd, J = 10.8, 5.4 Hz, 1H), 4.55-4.53 (m, 1H), 3.92 (d, J = 6.9

Hz, 1H), 3.35 (dd, J = 13.5, 5.4 Hz, 1H), 2.99-2.78 (m, 3H), 2.56-2.51 (m, 1H), 2.35-2.31 (m,

1H), 2.04-1.69 (m, 4H), 0.29 (s, 9H), 0.21 (s, 9H).

OHOH

(S)-65

reflux OP

OCl

(S)-66

toluene/THF, -40 °C → rt77% O

PO

N

(S,R,R)-64

Ph

Ph

PCl3

NPh

PhLi

(R,R)-68

O,O’-(S)-(1,1’-Dinaphthyl-2,2’-diyl)-N,N’-di-(R,R)-1-phenylethylphosphoramidite (64). In

the glove box, a 25 mL schlenk flask was charged with (S)-BINOL (65, 1.063 g, 3.714 mmol, 1

equiv) and PCl3 (4 mL, 46 mmol, 12 equiv). The flask was removed from the glove box and the

suspension was heated to reflux under Ar atmosphere. After refluxing 16 h, the mixture had

become a homogeneous solution. After an additional 8 h, the excess PCl3 was removed under

vacuum at 40 °C. The foamy residue was dissolved in toluene (3 mL) and the solvent was

44

removed under vacuum. This step was repeated two more times (2 × 3 mL toluene). The residue

remained under vacuum for 16 h after the final rinse. In a separate flask, n-BuLi (2.6 mL of a

1.6 M soln in hexanes, 4.2 mmol, 1.1 equiv) was added dropwise via syringe to a soln of bis((R)-

phenylethyl)amine (0.85 mL, 3.72 mmol, 1 equiv) in THF (13 mL) at -78 °C. The fuschia-

colored soln was allowed to warm to -40 °C. The phosphochloridite 66 was dissolved in toluene

(20 mL) and the soln was added to the deprotonated amine 68 via cannula. The amber-colored

soln was allowed to warm from -40 °C to rt and stirred 16 h. The resulting cloudy mixture was

filtered through a pad of celite and the filtrate was concentrated under reduced pressure. The

residue was purified by column chromatography using hexanes/EtOAc (14:1) as the eluent to

give pure phosphoramidite (S,R,R)-64 (1.770 g, 77%) as a white foamy solid. The 1H NMR

spectrum is in agreement with published values.96

O

MeO OMe

1 mol% Cu(OTf)22.4 mol% (S,R,R)-64

toluene, -25 °C50%

62

O

MeO OMe60

*

4,4-Dimethoxy-5-methyl-2-cyclohexenone (60). To a two-neck roundbottom flask fitted with a

thermometer adapter and containing a suspension of Cu(OTf)2 (0.059 g, 0.162 mmol, 0.010

equiv) in toluene (25 mL) was added a soln of the phosphoramidite ligand (S,R,R)-64 (0.201 g,

0.373 mmol, 0.024 equiv) in toluene (6 mL). The pale orange mixture stirred 1 h before being

cooled to -25 °C in a cryocool bath. The cyclohexadienone 6292 (2.390 g, 15.51 mmol, 1 equiv)

was added. Me2Zn (21 mL of a 1.2 M soln in toluene, 25 mmol, 1.6 equiv) was added so that the

internal temperature ≤-20 °C. The bright yellow reaction mixture stirred at -25 °C and the

reaction progress was monitored every 30 min by TLC. After 4 h, the slightly less polar

diaddition product was visible above the monoaddition product spot on the TLC plate. The

45

reaction was quenched by pouring the mixture into an Erlenmeyer flask containing rapidly

stirring 1.5 M NaOH. The aqueous layer was extracted with Et2O (3×). The combined organic

layers were washed with 1.5 N NaOH and brine, dried (Na2SO4), filtered and carefully

concentrated under reduced pressure to remove most of the Et2O. The remaining 3:1 mixture (by

1H NMR) of enone 60 and unreacted sm in toluene was applied to a silica gel column and eluted

with pentane/Et2O (3:1 to 0:1) to separate the toluene from the product mixture. Fractions

containing the product were combined and carefully concentrated under reduced pressure. The

residue was purified by column chromatography, eluting with pentane/Et2O (3:1) to give pure

cyclohexenone 60 (1.321 g, 50%) as a colorless oil. The 1H NMR spectrum is in agreement with

literature values.95

O

MeO OMe60

*

MgBr

THFMeO OMe

61

OH 1. BH3·DMS, then H2O, NaOH, H2O2

28

OH

2. TBSCl, DMAP, TEA3. Montmorillonite-K10

TBSO

O21%, 3 steps74%

4-(1-(tert-Butyldimethylsilyloxy)propan-2-yl)-4-hydroxy-6-methyl-2-cyclohexenone (28).

To the enone 60 (0.557 g, 3.28 mmol, 1 equiv) in THF (16 mL) at 0 °C was added

isopropenylmagnesium bromide (20 mL of a 0.5 M soln in THF, 10 mmol, 3 equiv) via syringe

pump over 20 min. The reaction was monitored by TLC and quenched by pouring the reaction

mixture into a flask containing saturated aqueous NH4Cl. The aqueous layer was extracted with

Et2O (4×). The combined organic portions were washed with brine, dried (MgSO4), filtered, and

concentrated under reduced pressure. The crude residue was flushed through a plug of pretreated

(TEA) silica gel with hexanes/EtOAc (4:1) to remove major impurities. Tertiary alcohol 61

(0.514 g, 74%) was obtained as an amber oil. 1H NMR (300 MHz, CDCl3): δ 5.80 (A of ABq, J

= 10.2 Hz, 1H), 5.74 (B of ABq, J = 10.2 Hz, 1H), 5.02 (s, 1H), 4.88 (quintet, J = 1.5 Hz, 1H),

46

3.28 (s, 3H), 3.22 (s, 3H), 2.18-2.12 (m, 2H), 1.77 (s, 3H), 1.68 (dd, J = 15.6, 6.6 Hz, 1H), 1.05

(d, J = 6.9 Hz, 3H).

To a soln of the olefin 61 (0.271 g, 1.276 mmol, 1 equiv) in THF (2 mL) was added

BH3·DMS (0.64 mL of a 2 M soln in THF, 1.28 mmol, 1 equiv) dropwise via syringe. The

reaction was monitored by 1H NMR analysis of aliquots. The reaction stirred 16 h, at which

point the flask was placed in an ambient temperature water bath and H2O (0.11 mL, 6.1 mmol, 5

equiv) was added followed by dropwise addition of NaOH (5 mL of a 3 M aqueous soln, 15

mmol, 12 equiv) and H2O2 (2.2 mL of a 33 wt% aqueous soln, 23.5 mmol, 18 equiv). The

mixture stirred 3 h and was diluted with Et2O. The aqueous layer was extracted with Et2O (4×).

The combined organic portions were washed with brine, dried (Na2SO4), decanted, and

concentrated under reduced pressure to give the crude primary alcohol (0.207 g, 71%).

To a soln of DMAP (11 mg, 0.09 mmol, 0.1 equiv) and the primary alcohol (0.207 g,

0.900 mmol, 1 equiv) in DCM (1 mL) was added TEA (0.31 mL, 2.22 mmol, 2.5 equiv)

followed by dropwise addition of a soln of TBSCl (0.163 g, 1.08 mmol, 1.2 equiv) in DCM (0.5

mL). Reaction progress was monitored by TLC. After 2 h, the mixture was diluted in H2O and

the aqueous layer was extracted with DCM (3×). The combined organic layers were washed

with brine, dried (Na2SO4), decanted, and concentrated under reduced pressure, giving the crude

silyl ether (0.310 g, 84%).

To a soln of the ketal (0.259 g, 0.752 mmol) in DCM (15 mL) was added

Montmorillonite K10 clay (0.378 g). Reaction progress was monitored by TLC. After 10 min,

additional clay was added (0.376 g). When no starting material was observable by TLC, the

reaction mixture was filtered through a pad of celite and the filtrate was concentrated under

reduced pressure. The crude residue, containing a 2:1 mix of diastereomers of 28 by 1H NMR,

47

was purified by column chromatography using 4:1 hexanes/EtOAc as the eluent to give the

desired enone 28 (0.082 g, 36% from 57). Partial separation of the diastereomers was achieved

during column chromatography. Major diastereomer: 1H NMR (300 MHz, CDCl3): δ 6.87 (dd, J

= 10.3, 2.1 Hz, 1H), 5.92 (d, J = 10.3 Hz, 1H), 4.43 (s, 1H), 4.07 (dd, J = 10.5, 3.3 Hz, 1H), 3.68

(dd, J = 10.5, 4.5 Hz, 1H), 2.48-2.36 (m, 2H), 2.08-1.94 (m, 2H), 1.17 (d, J = 6.3 Hz, 3H), 1.14

(d, J = 6.9 Hz, 3H), 0.93 (s, 9H), 0.13 (s, 6H); 13C NMR (75 MHz, CDCl3): δ 201.2, 154.7,

127.9, 73.6, 66.9, 43.7, 40.2, 38.8, 18.0, 15.8, 12.7, -5.7, -5.8; IR (neat, NaCl): 3463, 2931, 2858,

1684, 1462, 1377, 1073, 836 cm-1; MS m/z (%): 321 (12), 281 (16), 251 (80), 191 (13), 121 (29);

TOF-HRMS calcd for C16H30O3NaSi [M + Na]+ m/z: 321.1862, found: 321.1858; Minor

diastereomer: 1H NMR (300 MHz, CDCl3): δ 6.96 (dd, J = 10.2, 2.1 Hz, 1H), 5.92 (d, J = 10.2

Hz, 1H), 4.39 (s, 1H), 4.05 (dd, J = 10.2, 3.0 Hz, 1H), 3.69 (dd, J = 10.2, 2.7 Hz, 1H), 2.39 (dqd,

J = 13.5, 6.6, 4.7 Hz, 1H), 2.27 (ddd, J = 13.5, 4.7, 2.1 Hz, 1H), 1.99 (qt, J = 7.2, 3.0 Hz, 1H),

1.76 (t, J = 13.5 Hz, 1H), 1.19 (d, J = 7.2 Hz, 3H), 1.18 (d, J = 6.6 Hz, 3H), 0.94 (s, 9 H), 0.13 (s,

3H), 0.12 (s, 3H); 13C NMR (75 MHz, CDCl3): δ 201.3, 156.3, 127.3, 73.8, 67.1, 41.2, 39.1,

38.4, 25.8, 18.0, 15.7, 12.6, -5.7, -5.8.

48

APPENDIX A

1H AND 13C NMR SPECTRA FOR NEW COMPOUNDS

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

BIBLIOGRAPHY

1. Appendino, G.; Szallasi, A. Euphorbium: Modern Research on its Active Principle, Resiniferatoxin, Revives an Ancient Medicine. Life Sci. 1997, 60, 681-696.

2. Hergenhahn, M.; Adolf, W.; Hecker, E. Resiniferatoxin and Other Esters of Novel Polyfunctional Diterpenes from Euphorbia Resinifera and Unispina. Tetrahedron Lett. 1975, 1975, 1595-1598.

3. Adolf, W.; Sorg, B.; Hergenhahn, M.; Hecker, E. Structure-Activity Relations of Polyfunctional Diterpenes of the Daphnane Type. I. Revised Structure for Resiniferatoxin and Structure-Activity Relations of Resiniferonol and Some of Its Esters. J. Nat. Prod. 1982, 45, 347-354.

4. Szallasi, A.; Blumberg, P. M. Resiniferatoxin, a Phorbol-Related Diterpene, Acts as an Ultrapotent Analog of Capsaicin, the Irritant Constituent in Red Pepper. Neuroscience 1989, 30, 515-520.

5. Caterina, M. J.; Schumacher, M. A.; Tominaga, M.; Rosen, T. A.; Levine, J. D.; Julius, D. The Capsaicin Receptor: A Heat-Activated Ion Channel in the Pain Pathway. Nature 1997, 389, 816-824.

6. Chou, M. Z.; Mtui, T.; Gao, Y.-D.; Kohler, M.; Middleton, R. E. Resiniferatoxin Binds to the Capsaicin Receptor (TRPV1) Near the Extracellular Side of the S4 Transmembrane Domain. Biochemistry 2004, 43, 2501-2511.

7. Szallasi, A.; Blumberg, P. M. Vanilloid Receptors: New Insights Enhance Potential as a Therapeutic Target. Pain 1996, 68, 195-208.

8. Szallasi, A.; Blumberg, P. M. Vanilloid (Capsaicin) Receptors and Mechanisms. Pharmacological Rev. 1999, 51, 159-211.

9. Immke, D. C.; Gavva, N. R. The TRPV1 Receptor and Nociception. Semin. Cell Dev. Biol. 2006, 17, 582-591.

74

10. Cruz, F.; Guimaraes, M. Suppression of Bladder Hyperreflexia by Intravesical Resiniferatoxin. Lancet 1997, 350, 640-641.

11. Lazzeri, M.; Beneforti, P.; Turini, D. Urodynamic Effects of Intravesical Resiniferatoxin in Humans: Preliminary Results in Stable and Unstable Detrusor. J. Urol. 1997, 158, 2093-2096.

12. Payne, C. K.; Mosbaugh, P. G.; Forrest, J. B.; Evans, R. J.; Whitmore, K. E.; Antoci, J. P.; Perez-Marrero, R.; Jacoby, K.; Yu, A. S.; Frumkin, L. R. Intravesical Resiniferatoxin for the Treatment of Interstitial Cystitis: A Randomized, Double-Blind, Placebo Controlled Trial. J. Urol. 2005, 173, 1590-1594.

13. ICOS Corporation February 3, 2004 Press Release. http://media.corporate-ir.net/media_files/nsd/icos/news/PR_Q403_ICOS.pdf (October 2006).

14. Hanno, P.; Nordling, J.; van Ophoven, A. What is New in Bladder Pain Syndrome/Interstitial Cystitis? Curr. Opin. Urol. 2008, 18, 353-358.

15. Blumberg, P. M. Protein Kinase C as the Receptor for the Phorbol Ester Tumor Promoters: Sixth Rhoads Memorial Award Lecture. Cancer Res. 1988, 48, 1-8.

16. Wender, P. A.; Kee, J.-M.; Warrington, J. M. Practical Synthesis of Prostratin, DPP, and Their Analogs, Adjuvant Leads Against Latent HIV. Science 2008, 320, 649-652.

17. United States Food and Drug Administration List of Orphan Designations and Approvals. http://ww.fda.gov/orphan/designat/list.htm (November 2006).

18. Brown, D. C.; Iadarola, M. J.; Perkowski, S. Z.; Erin, H.; Shofer, F.; Laszlo, K. J.; Olah, Z.; Mannes, A. J. Physiologic and Antinociceptive Effects of Intrathecal Resiniferatoxin in a Canine Bone Cancer Model. Anesthesiology 2005, 103, 1052-1059.

19. Szallasi, A.; Cortright, D. N.; Blum, C. A.; Eid, S. R. The Vanilloid Receptor TRPV1: 10 Years From Channel Cloning to Antagonist Proof-of-Concept. Nat. Rev. Drug Discov. 2007, 6, 357-372.

20. Szallasi, A.; Cruz, F.; Geppetti, P. TRPV1: A Therapeutic Target for Novel Analgesic Drugs? Trends Mol. Med. 2006, 12, 545-554.

21. Kissin, I. Vanilloid-Induced Conduction Analgesia: Selective, Dose-Dependent, Long-Lasting, with a Low Level of Potential Neurotoxicity. Anesth. Analg. 2008, 107, 271-281.

22. Meyer, D. Use of Resiniferatoxin (RTX) for Producing an Agent for Treating Joint Pains and Method for Applying Said Agent. USPTO Application # 20080139641, 2008.

75

23. Fattorusso, E.; Lanzotti, V.; Taglialatela-Scafati, O.; Tron, G. C.; Appendino, G. Bisnorsesquiterpenoids from Euphorbia resinifera Berg. and an Expeditious Procedure to Obtain Resiniferatoxin from Its Fresh Latex. Eur. J. Org. Chem. 2002, 71-78.

24. Stanoeva, E.; He, W.; De Kimpe, N. Natural and Synthetic Cage Compounds Incorporating the 2,9,10-Trioxatricyclo[4.3.1.03,8]decane Type Moiety. Bioorg. Med. Chem. 2005, 13, 17-28.

25. Sakata, K.; Kawazu, K.; Mitsui, T.; Masaki, M. Structure and Stereochemistry of Huratoxin, a Piscicidal Constituent of Hura crepitans. Tetrahedron Lett. 1971, 1141-1144.

26. Yoshida, M.; Yokokura, H.; Hidaka, H.; Ikekawa, T.; Saijo, N. Mechanism of Antitumor Action of PKC Activator, Gnidimacrin. Int. J. Cancer 1998, 77, 243-250.

27. Kupchan, S. M.; Shizuri, Y.; Murae, T.; Sweeny, J. G.; Haynes, H. R.; Shen, M.-S.; Barrick, J. C.; Bryan, R. F.; van der Helm, D.; Wu, K. K. Gnidimacrin and Gnidimacrin 20-Palmitate, Novel Macrocyclic Antileukemic Diterpenoid Esters from Gnidia subcordata. J. Am. Chem. Soc. 1976, 98, 5719-5720.

28. Carney, J. R.; Krenisky, J. M.; Williamson, R. T.; Luo, J.; Carlson, T. J.; Hsu, V. L.; Moswa, J. L. Maprouneacin, a New Daphnane Diterpenoid with Potent Antihyperglycemic Activity from Maprounea africana. J. Nat. Prod. 1999, 62, 345-347.

29. He, W.; Cik, M.; Van Puyvelde, L.; Van Dun, J.; Appendino, G.; Lesage, A.; Van der Lindin, I.; Leysen, J. E.; Wouters, W.; Mathenge, S. G.; Mudida, F. P.; De Kimpe, N. Neurotrophic and Antileukemic Daphnane Diterpenoids from Synaptolepsis kirkii. Bioorg. Med. Chem. 2002, 10, 3245-3255.

30. Jayasuriya, H.; Zink, D. L.; Singh, S. B.; Borris, R. P.; Nanokorn, W.; Beck, H. T.; Balick, M. J.; Goetz, M. A.; Slayton, L.; Gregory, L.; Zakson-Aiken, M.; Shoop, W.; Singh, S. B. Structure and Stereochemistry of Rediocide A, a Highly Modified Daphnane from Trigonostemon reidioides Exhibiting Potent Insecticidal Activity. J. Am. Chem. Soc. 2000, 122, 4998-4999.

31. Zhan, Z.-J.; Fan, C.-Q.; Ding, J.; Yue, J.-M. Novel Diterpenoids with Potent Inhibitory Activity Against Endothelium Cell HMEC and Cytotoxic Activities from a Well-known TCM Plant Daphne genkwa. Bioorg. Med. Chem. 2005, 13, 645-655.

32. Wender, P. A.; Jesudason, C. D.; Nakahira, H.; Tamura, N.; Tebbe, A. L.; Ueno, Y. The First Synthesis of a Daphnane Diterpene: The Enantiocontrolled Total Synthesis of (+)-Resiniferatoxin. J. Am. Chem. Soc. 1997, 119, 12976-12977.

76

33. Jackson, S. R.; Johnson, M. G.; Mikami, M.; Shiokawa, S.; Carreira, E. M. Rearrangement of a Tricyclic 2,5-Cyclohexadienone: Towards a General Synthetic Route to the Daphnanes and (+)-Resiniferatoxin. Angew. Chem. Int. Ed. 2001, 40, 2694-2697.

34. Ritter, T.; Zarotti, P.; Carreira, E. M. Diastereoselective Phenol para-Alkylation: Access to a Cross-Conjugated Cyclohexadienone en Route to Resiniferatoxin. Org. Lett. 2004, 6, 4371-4374.

35. Wender, P. A.; Kogen, H.; Lee, H. Y.; Munger, J. D., Jr.; Wilhelm, R. S.; Williams, P. D. Studies on Tumor Promoters. 8. The Synthesis of Phorbol. J. Am. Chem. Soc. 1989, 111, 8957-8958.

36. Wender, P. A.; Lee, H. Y.; Wilhelm, R. S.; Williams, P. D. Studies on Tumor Promoters. 7. The Synthesis of a Potentially General Precursor of the Tiglianes, Daphnanes, and Ingenanes. J. Am. Chem. Soc. 1989, 111, 8954-8957.

37. Wender, P. A.; McDonald, F. E. Studies on Tumor Promoters. 9. A Second-Generation Synthesis of Phorbol. J. Am. Chem. Soc. 1990, 112, 4956-4958.

38. Wender, P. A.; Rice, K. D.; Schnute, M. E. The First Formal Asymmetric Synthesis of Phorbol. J. Am. Chem. Soc. 1997, 119, 7897-7898.

39. Wender, P. A.; D'Angelo, N.; Elitzin, V. I.; Ernst, M.; Jackson-Ugueto, E. E.; Kowalski, J. A.; McKendry, S.; Rehfeuter, M.; Sun, R.; Voigtlaender, D. Function-Oriented Synthesis: Studies Aimed at the Synthesis and Mode of Action of 1α-Alkyldaphnane Analogues. Org. Lett. 2007, 9, 1829-1832.

40. Lee, K.; Cha, J. K. Formal Synthesis of (+)-Phorbol. J. Am. Chem. Soc. 2001, 123, 5590-5591.

41. Brickwood, A. C.; Drew, M. G. B.; Harwood, L. M.; Ishikawa, T.; Marais, P.; Morisson, V. Synthetic Approaches Towards Phorbols via the Ultra-High-Pressure Mediated Intramolecular Diels-Alder Reaction of Furans (IMDAF): Effect of Furan Substitution. J. Chem. Soc., Perkin Trans. 1 1999, 913-921.

42. Page, P. C. B.; Hayman, C. M.; McFarland, H. L.; Willcock, D. J.; Galea, N. M. An IMDA Approach to Tigliane and Daphnane Diterpenoids: Generation of Rings A, B and C Incorporating C-18. Synlett 2002, 583-587.

43. Rigby, J. H.; Kierkus, P. C.; Head, D. Studies on the Stereoselective Construction of the Tigliane Ring System. Tetrahedron Lett. 1989, 30, 5073-5076.

77

44. Shigeno, K.; Sasai, H.; Shibasaki, M. Synthetic Studies Towards Phorbols: A Stereocontrolled Synthesis of the Phorbol Skeleton in the Naturally Occurring Form. Tetrahedron Lett. 1992, 33, 4937-4940.

45. Sugita, K.; Shigeno, K.; Neville, C. F.; Sasai, H.; Shibasaki, M. Synthetic Studies Towards Phorbols: Synthesis of B or C Ring Substituted Phorbol Skeletons in the Naturally Occurring Form. Synlett 1994, 325-329.

46. McLoughlin, J. I.; Brahma, R.; Campopiano, O.; Little, R. D. Stereoselectivity in Intramolecular Diyl Trapping Reactions. Model Studies Directed Toward the Phorbols. Tetrahedron Lett. 1990, 31, 1377-1380.

47. Paquette, L. A.; Sauer, D. R.; Edmondson, S. D.; Friedrich, D. A Concise Route to the Tetracyclic Core of Phorbol. Tetrahedron Lett. 1994, 50, 4071-4086.

48. Brummond, K. M.; Chen, D.; Davis, M. M. A General Synthetic Route to Differentially Functionalized Angularly and Linearly Fused [6-7-5] Ring Systems: A Rh(I)-Catalyzed Cyclocarbonylation Reaction. J. Org. Chem. 2008, 73, 5064-5068.

49. Brummond, K. M.; Chen, H.; Fisher, K. D.; Kerekes, A. D.; Rickards, B.; Sill, P. C.; Geib, S. J. An Allenic Pauson-Khand-Type Reaction: A Reversal in π-Bond Selectivity and the Formation of Seven-Membered Rings. Org. Lett. 2002, 4, 1931-1934.

50. Brummond, K. M.; Gao, D. Unique Strategy for the Assembly of the Carbon Skeleton of Guanacastapene A Using an Allenic Pauson-Khand-Type Reaction. Org. Lett. 2003, 5, 3491-3494.

51. Brummond, K. M.; Lu, J. A Short Synthesis of the Potent Antitumor Agent (±)-Hydroxymethylacylfulvene Using an Allenic Pauson-Khand Type Cycloaddition. J. Am. Chem. Soc. 1999, 121, 5087-5088.

52. Brummond, K. M.; Lu, J.; Petersen, J. A Rapid Synthesis of Hydroxymethylacylfulvene (HMAF) Using the Allenic Pauson-Khand Reaction. A Synthetic Approach to Either Enantiomer of This Illudane Structure. J. Am. Chem. Soc. 2000, 122, 4915-4920.

53. Brummond, K. M.; Wan, H. The Allenic Pauson-Khand Cycloaddition. Dependence on π-Bond Selectivity on Substrate Structure. Tetrahedron Lett. 1998, 39, 931-934.

54. Brummond, K. M.; Wan, H.; Kent, J. L. An Intramolecular Allenic [2 + 2 + 1] Cycloaddition. J. Org. Chem. 1998, 63, 6535-6545.

78

55. Kent, J. L.; Wan, H.; Brummond, K. M. A New Allenic Pauson-Khand Cycloaddition for the Preparation of α-Methylene Cyclopentenones. Tetrahedron Lett. 1995, 36, 2407-2410.

56. Curran, D. P. The Cycloadditive Approach to β-Hydroxy Carbonyls: An Emerging Alternative to the Aldol Strategy. In Advances in Cycloaddition; Curran, D. P., Ed.; JAI Press Inc.: Greenwich, CT, 1988; Vol. 1, pp 129-189.

57. Jäger, V.; Colinas, P. Nitrile Oxides. In Synthetic Applications of 1,3-Dipolar Cycloaddition Chemistry Toward Heterocycles and Natural Products; Padwa, A., Pearson, W. H., Eds.; John Wiley & Sons: New York, 2002; pp 361-472.

58. Belen'kii, L. I. Nitrile Oxides. In Nitrile Oxides, Nitrones, and Nitronates in Organic Synthesis: Novel Strategies in Synthesis; 2nd ed.; Feuer, H., Ed.; John Wiley & Sons: Hoboken, NJ, 2008; pp 1-127.

59. Caramella, P.; Grünanger, P. Nitrile Oxides and Imines. In 1,3-Dipolar Cycloaddition Chemistry; Padwa, A., Ed.; John Wiley & Sons: New York, 1984; Vol. 1, pp 291-392.

60. Grundmann, C.; Grünanger, P. The Nitrile Oxides: Versatile Tools of Theoretical and Preparative Chemistry,Springer-Verlag: New York, 1971.

61. Torssell, K. B. G. Nitrile Oxides, Nitrones, and Nitronates in Organic Synthesis: Novel Strategies in Synthesis,1st ed.; VCH Publishers, Inc.: New York, 1988.

62. Bianchi, G.; De Micheli, C.; Gandolfi, R.; Grünanger, P.; Finzi, P. V.; de Pava, O. V. Isoxazoline Derivatives. Part VI. Regioselectivity in the 1,3-Dipolar Cycloaddition of Nitrile Oxides to α,β-Unsaturated Ketones. J. Chem. Soc., Perkin Trans. 1 1973, 1148-1155.

63. Caramella, P.; Cellerino, G. Selectivity in Cycloadditions - II. Polar and Steric Control in the 1,3-Dipolar Cycloaddition of Benzonitrile Oxide to Some 3-Substituted Cyclopentenes. Tetrahedron Lett. 1974, 229-232.

64. Houk, K. N.; Moses, S. R.; Wu, Y.-D.; Rondan, N. G.; Jäger, V.; Schohe, R.; Fronczek, F. R. Stereoselective Nitrile Oxide Cycloadditions to Chiral Allyl Ethers and Alcohols. The "Inside Alkoxy" Effect. J. Am. Chem. Soc. 1984, 106, 3880-3882.

65. Curran, D. P.; Choi, S.-M.; Gothe, S. A.; Lin, F.-T. Directed Nitrile Oxide Cycloaddition Reactions. The Use of Hydrogen Bonding to Direct Regio- and Stereochemistry in Nitrile Oxide Cycloadditions with Cyclopentenylamides. J. Org. Chem. 1990, 55, 3710-3712.

79

66. Becker, N.; Carreira, E. M. Hydroxyl-Directed Nitrile Oxide Cycloaddition Reactions with Cyclic Allylic Alcohols. Org. Lett. 2007, 9, 3857-3858.

67. Bode, J. W.; Fraefel, N.; Muri, D.; Carreira, E. M. A General Solution to the Modular Synthesis of Polyketide Building Blocks by Kanemasa Hydroxy-Directed Nitrile Oxide Cycloadditions. Angew. Chem. Int. Ed. 2001, 40, 2082-2085.

68. Choi, S.-M. Part 1. Hydrogen-Bond Directing Nitrile Oxide Cycloadditions Part II. 1,2-Asymmetric Induction by Radical and Ionic Reductions. Ph.D. Thesis, University of Pittsburgh, 1992.

69. Kim, H. R.; Song, J. H.; Rhie, S. Y.; Ryu, E. K. Regioselective and Stereoselective 1,3-Dipolar Cycloadditions of Nitrile Oxide with Allylic Alcohols Prepared in Situ from α,β-Unsaturated Carbonyl Compounds with Grignard Reagents. Synth. Commun. 1995, 25, 1801-1807.

70. Kanemasa, S.; Nishiuchi, M.; Kamimura, A.; Hori, K. First Successful Metal Coordination Control in 1,3-Dipolar Cycloadditions. High-Rate Acceleration and Regio- and Stereocontrol of Nitrile Oxide Cycloadditions to the Magnesium Alkoxides of Allylic and Homoallylic Alcohols. J. Am. Chem. Soc. 1994, 116, 2324-2339.

71. Martin, S. F.; Anderson, B. G.; Daniel, D.; Gaucher, A. A Cycloaddition Approach to Breynolide. Tetrahedron 1997, 53, 8997-9006.

72. Martin, S. F.; Daniel, D. A Novel Approach to Breynolide. Tetrahedron Lett. 1993, 34, 4281-4284.

73. Seifried, D. unpublished results.

74. Mukaiyama, T.; Hoshino, T. The Reactions of Primary Nitroparaffins with Isocyanates. J. Am. Chem. Soc. 1960, 82, 5339-5342.

75. Kantorowski, E. J.; Brown, S. P.; Kurth, M. J. Use of Diisocyanates for in Situ Preparation of Nitrile Oxides: Preparation of Isoxazoles and Isoxazolines. J. Org. Chem. 1998, 63, 5272-5274.

76. 3.7-6.6 Hz is a typical 3JH4,H5 coupling for trans-isoxazolines; see ref 77.

77. Aversa, M. C.; Cum, G.; Crisafulli, M. Spettri NMR di Δ2-Isossazoline. - Nota II. Isomeria cis-trans di Δ2-Isossazoline 4,5-Bisostituite. Gazz. Chim. Ital. 1968, 98, 42-47.

78. Chiu, P.; Li, Z.; Fung, K. C. M. An Expedient Preparation of Stryker's Reagent. Tetrahedron Lett. 2003, 44, 455-457.

80

79. Kamenecka, T. M.; Overman, L. E.; Ly Sakata, S. K. Construction of Substituted Cyclohexenones by Reductive Cyclization of 7-Oxo-2,8-alkadienyl Esters. Org. Lett. 2002, 4, 79-82.

80. Lipshutz, B. H.; Chrisman, W.; Noson, K.; Papa, P.; Sclafani, J. A.; Vivian, R. W.; Keith, J. M. Copper Hydride-Catalyzed Tandem 1,4-Reduction/Alkylation Reactions. Tetrahedron 2000, 56, 2779-2788.

81. Lipshutz observes significant quantities of impurities in two samples of Stryker's reagent obtained frou Aldrich (see ref 80). We generally purchased our commercial Stryker's reagent from Acros. The quality of Stryker's reagent can be judged to some degree by the appearance of the copper hydride. Freshly prepared Stryker's reagent is bright red to deep red and crystalline, whereas quenched Stryker's reagent is dark brown. The typical sample we received from Acros ranged from brick red to brownish-red and often appeared somewhat heterogeneous.

82. Lee, D.-W.; Yun, J. Direct Synthesis of Stryker's Reagent From a Cu(II) Salt. Tetrahedron Lett. 2005, 46, 2037-2039.

83. Caldirola, P.; De Amici, M.; De Micheli, C.; Wade, P. A.; Price, D. T.; Bereznak, J. F. Metal-Hydride Reduction of Isoxazoline-3-Carboxylate Esters. Tetrahedron 1986, 42, 5267-5272.

84. Kamimura, A.; Kakehi, A.; Hori, K. An Experimental and Theoretical Study on Stereoselective Addition to 3-Formyl-Δ2-isoxazolines. Part 1. 1,3-Antiselectivity Induced by BF3�OEt2. Tetrahedron 1993, 49, 7637-7648.

85. Yamamoto, H. Propargyl and Allenyl Organometallics. In Comprehensive Organic Synthesis; Trost, B., Fleming, I., Eds.; Pergamon Press: Elmsford, NY, 1991; Vol. 2, pp 81-98.

86. Daniels, R. G.; Paquette, L. A. Silanes in Organic Synthesis. II. Regiocontrolled Synthesis of α-Hydroxymethylated (Trimethylsilyl)allenes. Tetrahedron Lett. 1981, 22, 1579-1582.

87. Daniels, R. G. Regiochemical and Stereochemical Studies of Organosilanes. Ph.D. Thesis, Ohio State University, Columbus, OH, 1982.

88. Lin, M.-J.; Loh, T.-P. Indium-Mediated Reaction of Trialkylsilyl Propargyl Bromide with Aldehydes: Highly Regioselective Synthesis of Allenic and Homopropargylic Alcohols. J. Am. Chem. Soc. 2003, 125, 13042-13043.

81

89. Cabezas, J. A.; Alvarez, L. X. Propargylation of Carbonyl Compounds: An Efficient Method for the Synthesis of Homopropargyl Alcohols. Tetrahedron Lett. 1998, 39, 3935-3938.

90. Cabezas, J. A.; Pereira, A. R.; Amey, A. A New Method for the Preparation of 1,3-Dilithiopropyne: An Efficient Synthesis of Homopropargyl Alcohols. Tetrahedron Lett. 2001, 42, 6819-6822.

91. Gotō, H.; Ōsawa, E. An Efficient Algorithm for Searching Low-energy Conformers of Cyclic and Acyclic Molecules. J. Chem. Soc., Perkin Trans. 2 1993, 187-198.

92. Pelter, A.; Elgendy, S. M. A. Phenolic Oxidations with Phenyliodonium Diacetate. J. Chem. Soc., Perkin Trans. 1 1993, 1891-1896.

93. Nilsson, A.; Ronlán, A. A Novel Synthesis of 4-Chloro-4-Methylcyclohexa-2,5-dienone and 4,4-Dimethoxycyclohexa-2,5-dienone. Tetrahedron Lett. 1975, 1107-1110.

94. Stern, A. J.; Rohde, J. J.; Swenton, J. S. Oxygenophilic Organoaluminum-Mediated Conjugate Addition of Alkyllithium and Grignard Reagents to Quinone Monoketals and Quinol Ethers. The Directing Effect of a Methoxy Group on the 1,4-Addition Process. J. Org. Chem. 1989, 54, 4413-4419.

95. Imbos, R.; Brilman, M. H. G.; Pineschi, M.; Feringa, B. L. Highly Enantioselective Catalytic Conjugate Additions to Cyclohexadienones. Org. Lett. 1999, 1, 623-625.

96. Arnold, L. A.; Imbos, R.; Mandoli, A.; de Vries, A. H. M.; Naasz, R.; Feringa, B. L. Enantioselective Catalytic Conjugate Addition of Dialkylzinc Reagents Using Copper-Phosphoramidite Complexes; Ligand Variation and Non-linear Effects. Tetrahedron 2000, 56, 2865-2878.

97. Imbos, R. Catalytic Asymmetric Conjugate Additions and Heck Reactions. Ph.D. Thesis, University of Groningen, Groningen, Netherlands, 2002.

98. The Feringa group has commented on the somewhat variable purity of phosphoramidite ligands. See ref 97.

99. Duursma, A. Asymmetric Catalysis with Chiral Monodentate Phosphoramidite Ligands. Ph.D. Thesis, University of Groningen, Groningen, Netherlands, 2004.

100. The identities of the ligand impurities are unknown, although evidence of their existence can be seen by both 1H and 31P NMR. Extraneous peaks are visible in the olefinic region of the 1H NMR spectrum of impure ligand. In addition to the expected resonance at δ 146.0 ppm in the 31P NMR, impure samples of phosphoramidite ligand 64 also gave

82

83

small peaks at δ 139.7 and 138.1 ppm. The impurities are not visible by TLC and were not removed by either column chromatography or recrystallization.

101. Ley, S. V.; Anthony, N. J.; Armstrong, A.; Brasca, M. G.; Clarke, T.; Culshaw, D.; Greck, C.; Grice, P.; Jones, A. B.; Lygo, B.; Madin, A.; Sheppard, R. N.; Slawin, A. M. Z.; Williams, D. J. A Highly Convergent Total Synthesis of the Spiroacetal Macrolide (+)-Milbemycin β1. Tetrahedron 1989, 45, 7161-7194.

102. Paquette, L. A.; Earle, M. J.; Smith, G. F. (4R)-(+)-tert-Butyldimethylsiloxy-2-cyclopenten-1-one. In Organic Syntheses; Wiley & Sons: New York, 1998; Collect. Vol. No. 9, pp 132-136.

103. Gautier, E. C. L.; Graham, A. E.; McKillop, A.; Standen, S. P.; Taylor, R. J. K. Acetal and ketal Deprotection using Montmorillonite K10: The First Synthesis of syn-4,8-Dioxatricyclo[5.1.0.03,5]-2,6-octanedione. Tetrahedron Lett. 1997, 38, 1881-1884.

104. Moore, J. E.; Davies, M. W.; Goodenough, K. M.; Wybrow, R. A. J.; York, M.; Johnson, C. N.; Harrity, J. P. A. Investigation of the Scope of a [3 + 2] Cycloaddition Approach to Isoxazole Boronic Esters. Tetrahedron 2005, 61, 6707-6714.

105. Kim, J. N.; Ryu, E. K. 1,3-Dipolar Cycloaddition: Molecular Sieves Assisted Generation of Nitrile Oxides from Hydroximoyl Chlorides. Heterocycles 1990, 31, 1693-1697.

106. Still, W. C.; Kahn, M.; Mitra, A. Rapid Chromatographic Technique for Preparative Separations with Moderate Resolution. J. Org. Chem. 1978, 43, 2923-2925.